Next Article in Journal
Surface Chemical Effects on Fischer–Tropsch Iron Oxide Catalysts Caused by Alkali Ion (Li, Na, K, Cs) Doping
Previous Article in Journal
High-Performance Methanol Oxidation via Ni12-Metal8/CNF Catalyst for Fuel Cell Applications
Previous Article in Special Issue
Impact of Inorganic Anions on the Photodegradation of Herbicide Residues in Water by UV/Persulfate-Based Advanced Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interstitial N-Doped TiO2 for Photocatalytic Methylene Blue Degradation under Visible Light Irradiation

1
School of Chemical and Environmental Engineering, Liaoning University of Technology, Jinzhou 121001, China
2
Departament de Química, Facultat de Ciències, Universitat Autònoma de Barcelona (UAB), Cerdanyola del Valles, 08193 Barcelona, Spain
*
Authors to whom correspondence should be addressed.
These authors contributed equally to the manuscript.
Catalysts 2024, 14(10), 681; https://doi.org/10.3390/catal14100681
Submission received: 20 August 2024 / Revised: 6 September 2024 / Accepted: 23 September 2024 / Published: 1 October 2024
(This article belongs to the Special Issue Recent Advances in Photocatalytic Treatment of Pollutants in Water)

Abstract

:
Photocatalysis is a promising method for methylene blue (MB) degradation due to its effectiveness and environmental compatibility. Among the photocatalysts, titanium dioxide (TiO2) has been widely used for MB degradation due to its exceptional photocatalytic activity. However, the wide bandgap limits the degradation efficiency of TiO2 under visible light. Here, an interstitial nitrogen-doped TiO2 (5%NT/TiO2) used thiourea as the N source was fabricated for visible light-derived MB degradation. The 5%NT/TiO2 exhibited an extended absorption range of visible light. Moreover, photoelectrochemical measurements showed an improvement in the photocurrent response and charge transfer behavior on N/TiO2. Thus, 5%NT/TiO2 had enhanced photocatalytic activity compared with pristine TiO2 and substitutive N-doped TiO2 (5%NAB/TiO2). The accelerated photocatalytic MB degradation process on N/TiO2 could be mainly attributed to the interstitial N doping, which caused the appearance of new energy states and extended optical properties. Through comparing the impact of interstitial and substitutive in TiO2 activity, our work proposes a suitable form of element doping to enhance the optical properties and photocatalytic activity of TiO2 and even other semiconductors, providing guidance for future work.

Graphical Abstract

1. Introduction

Methylene blue (MB) is a common dye used in various industries [1,2]. However, with its widespread utilization, MB leads to significant environmental challenges, including water pollution, toxicity to aquatic life, bioaccumulation, and potential human health risks [3,4]. Thus, effective removal of MB is crucial for ecosystems and human health. Several strategies have been developed for MB remediation, such as advanced oxidation processes (AOPs), adsorption techniques, bioremediation, and membrane filtration [5,6]. Among them, based on AOP technology, photocatalytic degradation of organic compounds is an ideal method for the removal of industrial pollutants with solar energy as the only input energy source. Improving photocatalytic degradation efficiency has been attracting a great deal of research interest.
The design of the photocatalyst is so important for photocatalytic degradation that it determines the whole reaction’s efficiency [7,8,9,10]. Titanium dioxide (TiO2) is a widely studied semiconductor material known for its excellent photocatalytic properties, chemical stability, and non-toxicity. It has garnered significant attention for applications in environmental remediation, particularly in the degradation of organic pollutants. However, the practical application of TiO2 is limited by its wide bandgap (~3.2 eV for anatase), which restricts its photo-response to the ultraviolet light region, constituting only about 5% of the solar spectrum. In order to improve the photoconversion efficiency and photocatalytic degradation properties of TiO2, it is necessary to develop a reliable strategy for extending the photo-response range of TiO2 photocatalysts. Nitrogen doping is a promising approach to modify the electronic structure of TiO2 and enhance its capabilities for visible light absorption [11,12]. By incorporating nitrogen atoms into the TiO2 lattice, the bandgap can be narrowed, allowing it to absorb visible light and thus improving its photocatalytic efficiency under solar irradiation [13,14,15]. Although numerous studies have proved the positive influences of N element doping in the photocatalytic activity of TiO2, the effects of different N precursors on achieving N doping should be further explored to elucidate the impact of N atoms on the structure and activity of TiO2.
In this study, nitrogen-doped TiO2 catalysts (denoted x% NT/TiO2 and x% NAB/TiO2, with x being the wt% of N element) were prepared from two different N sources (i.e., thiourea and ammonium bicarbonate) and used for photocatalytic MB degradation under visible light irradiation. Compared with TiO2 and NAB/TiO2, NT/TiO2 showed better photocatalytic degradation performance. The content of N elements was optimized, showing that 5% NT/TiO2 possessed the highest MB degradation efficiency, with more than 60% of the MB removed within 150 min. In order to elucidate the effect of N element doping and clarify the advantages of thiourea as the N source, a series of characterizations were conducted, which confirmed the interstitial N doping in NT/TiO2. N atoms from thiourea were mainly incorporated into the lattice of NT/TiO2. On the other hand, N atoms from ammonium bicarbonate were mainly incorporated into NAB/TiO2 in the form of oxygen substitution. Moreover, NT/TiO2 possessed an improved capacity for visible light absorption. The photoelectrochemical experiments indicated that the interstitial N doping increased the number of photoexcited electrons from NT/TiO2 under visible light illumination. Therefore, with enhanced optical properties and photoelectric response, NT/TiO2 with interstitial N atoms displayed the optimal MB degradation efficiency. This work proved the advantages of interstitial N doping in improving the photocatalytic activity of semiconductors represented by TiO2.

2. Results and Discussion

2.1. Optimization of the Amount of N Doping

The concentration of nitrogen doping was optimized by gradually increasing the amount of nitrogen elements in 5%NT/TiO2 in terms of the degradation efficiency of MB. Figure 1 shows the efficiencies of photocatalytic MB degradation with different amounts of N doping. Under visible light irradiation, undoped TiO2 exhibited ~15% MB degradation after 150 min (Figure 1). The 5%NT/TiO2 had the highest MB photodegradation efficiency. After 150 min, approximately 56% of the MB was degraded. The decline in the activity of 10%NT/TiO2 and 15%NT/TiO2 could be attributed to the excessive N atoms, which may be recombination sites of photoexcited charges, as found by previous reports [16,17,18]. After determining the optimal amount of N doping, namely 5%, in the subsequent research and analysis, the amount of doping remained consistent, and only the N source was changed.

2.2. XRD Patterns of TiO2, 5%NT/TiO2, and 5%NAB/TiO2 Catalysts

X-ray diffraction (XRD) analysis was performed to determine the crystalline phases and crystallographic structure of 5%NT/TiO2, 5%NT/TiO2, and TiO2 catalysts (Figure 2a). The undoped TiO2 mainly consisted of the rutile crystalline phase. The peaks centered at 27.5°, 36.1°, 41.3°, and 54.3° corresponded to the (110), (101), (111), and (211) crystal planes of the rutile phase, respectively [19]. When TiO2 was doped with thiourea as the nitrogen source, 5%NT/TiO2 mainly showed the anatase phase. The peaks centered at 25.2°, 37.8°, 48.0°, and 55.0° corresponded to the (101), (004), (200), and (211) crystal planes of the anatase phase, respectively [19]. However, the 5%NAB/TiO2 with ammonium bicarbonate as the nitrogen source mainly showed a mixed crystal phase of rutile and anatase. This transformation in the crystal form may be attributed to the fact that when N is doped into the TiO2 lattice, it will replace some of the O atoms or enter the lattice’s interstitial positions. This will cause a distortion in the lattice structure of TiO2 and generate stress. When the stress accumulates to a certain extent, in order to reduce the system’s energy, the crystal structure will change, thus leading to a change in the crystal form. Moreover, the (101) crystal plane of 5%NT/TiO2 shifted from 25.2° to 26.1° (Figure 2b). It indicated that the interstitial doping, which can cause lattice distortion, had been achieved in 5%NT/TiO2. However, 5%NAB/TiO2 maintained two TiO2 phases without a peak shift, meaning that the doping mode of N in 5%NAB/TiO2 may be substitutive doping.

2.3. SEM Images and Corresponding Element Mappings of 5%NT/TiO2

The elemental distribution of the 5%NT/TiO2 catalyst was observed by scanning electron microscopy (SEM). As can be seen from Figure 3b–d, the N element showed a very uniform distribution state on the sample and had the same distribution as the Ti and O elements. This phenomenon strongly indicated that the N element had been successfully doped into the lattice of TiO2.

2.4. XPS Spectra of TiO2, 5%NT/TiO2, and 5%NAB/TiO2 Catalysts

X-ray photoelectron spectroscopy (XPS) analysis was conducted to investigate the surface elemental composition of the TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts. The results revealed that the XPS results of the catalysts showed the presence of Ti 2p, O 1s, and N 1s peaks (Figure 4). The XPS spectra of 5%NT/TiO2 and 5%NAB/TiO2 revealed the presence of nitrogen, indicating that nitrogen was doped into TiO2.
Figure 4a illustrates the HR-XPS spectra of Ti 2p in TiO2, 5%NT/TiO2, and 5%NAB/TiO2. For TiO2, the Ti 2p3/2 and 2p1/2 core energy level peaks appeared at 459.1 and 464.9 eV, respectively, which were contributed by the O–Ti–O in TiO2 [20,21,22]. Compared with TiO2, in 5%NT/TiO2, the Ti 2p3/2 and 2p1/2 core energy level peaks appeared at 458.8 and 464.5 eV, respectively, which were the Ti 2p peaks of the N–Ti–N or O–Ti–N in N–TiO2 [21,22,23]. For 5%NAB/TiO2, the Ti 2p peaks of 5%NAB/TiO2 shifted to the lower binding energy with changes of ~0.4 eV. The peak shifts of Ti 2p indicated the successful incorporation of N in both 5%NT/TiO2 and 5%NAB/TiO2 [20,21].
In Figure 4b, TiO2 exhibits an O 1s peak at 529.9 eV, which was attributed to the lattice oxygen (Ti–O–Ti). And the peak at 531.8 eV was surface-adsorbed oxygen. In 5%NAB/TiO2 and 5%NT/TiO2, the peaks had an obvious shift owing to the interstitial doping of N into the TiO2 lattice to form hyponitrite (N2O2)2–, which was further confirmed in the FTIR analyses, as discussed below.
Figure 4c clearly shows the binding energies of N 1s at 399.8 and 401.9 eV for 5%NT/TiO2 and 5%NAB/TiO2, respectively. For the 5%NT/TiO2, according to previous literature [24], the peak at 401.9 eV indicated the formation of Ti–O–N bonds based on the form of interstitial N in the lattice gap of 5%NT/TiO2. In the XPS spectrum of 5%NAB/TiO2, the peak of N 1s at 399.8 eV could be attributed to the Ti–N–Ti bonding, which was due to the substitution of oxygen in the 5%NT/TiO2 lattice by N and thus was consistent with the XRD results.
Through a comparison of the XPS spectra, the difference between using thiourea and ammonium bicarbonate as N source could be attributed to the different chemical bonds between N and TiO2. In NT/TiO2, more N atoms existed in the form of interstitial N in the lattice gap of TiO2 and formed Ti–O–N bonds, indicating that the O atoms had not been replaced. The following Fourier transform infrared spectroscopy (FTIR) spectra also confirmed this idea due to the existence of N+–O types such as (N2O2)2–. However, in NAB/TiO2, more N atoms tended to form Ti–N–Ti bonds, suggesting that many O atoms were replaced by N atoms.

2.5. TEM Images of TiO2, 5%NT/TiO2, and 5%NAB/TiO2 Catalysts

Transmission electron microscopy (TEM) characterization was used to observe the morphology of the TiO2 and N/TiO2 catalysts. Undoped TiO2 typically showed mainly rutile phases with particle sizes ranging from 25 to 50 nm (Figure 5a). After N element doping, the crystal phase of 5%NT/TiO2 transformed into the anatase phase, while 5%NAB/TiO2 became mainly a mixture of rutile and anatase phases (Figure 5b,c). The particle sizes of 5%NT/TiO2 and 5%NAB/TiO2 were slightly smaller than that of TiO2. Among them, 5%NT/TiO2 had the smallest particle size (5.5–8 nm). High-resolution transmission eectron microscopy (HRTEM) images showed clear lattice fringes with an interplanar spacing of 3.20 Å and 2.48 Å (Figure 5d,f), corresponding to the (110) and (101) facets of rutile TiO2, respectively. The lattice spacings of 3.48 Å and 3.51 Å (Figure 5e,f), respectively, corresponded to those of the (101) plane of anatase TiO2 in 5%NT/TiO2 and 5%NAB/TiO2. It can be observed that due to the difference in the doping methods, significant differences in the lattice spacings of 5%NT/TiO2 and 5%NAB/TiO2 were induced. This was because 5%NT/TiO2 involved interstitial doping, and the N doping was in the lattice gap of TiO2, which led to lattice distortion.

2.6. N2 Adsorption–Desorption Isotherm Analysis of TiO2, 5%NT/TiO2, and 5%NAB/TiO2 Catalysts

N2 adsorption–desorption isotherms analysis was used to determine the specific surface area and porosity of the TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts. The N2 isotherms of TiO2, 5%NT/TiO2, and 5%NAB/TiO2 displayed IV-type isotherms (Figure 6a). This indicated that all the catalysts had mesoporous structures with pore size distributions of 2–40 nm (Figure 6b). The 5%NT/TiO2 had a lower pore size distribution. Figure 6b and Table 1 show that the N doping, which can create additional surface defects and porosity, led to higher specific surface area and pore size. The larger specific surface area is beneficial for mass transfer during photocatalytic reactions. According to previous reports, the N atoms can affect the properties of porosity, including the average pore size and distribution, and the specific surface area [25,26]. In detail, under high-temperature processing, NH3 gas was produced via in-situ breakdown of ammonia precursors (e.g., thiourea) so that the porous structure was formed during synthesis [27,28]. On the other hand, the decomposition of ammonium bicarbonate produced CO2, further leading to the formation of interconnected mesopores in the TiO2 lattice [29].

2.7. FTIR of TiO2, 5%NT/TiO2, and 5%NAB/TiO2 Catalysts

FTIR spectroscopy was employed to identify the presence of nitrogen-related species. (Figure 7). The characteristic peaks in the low wavenumber range (500 cm−1) were attributed to TiO2. The spectra showed characteristic peaks of TiO2 at 3400 cm−1, 1630 cm−1, and 700–500 cm−1, which were attributed to the Ti–OH bond, the OH bending vibration of water molecules, and Ti–O–Ti bond stretching vibrations, respectively. Notably, the 5%NT/TiO2 presented an additional peak at 1388 cm−1, which was also attributed to the stretching vibration of N–O. Several bands in the low-frequency region observed at 1050 and 1250 cm−1 belonged to the N+–O type substances embedded in the TiO2 network. The peak at 1150 cm−1 was the vibration peak of (N2O2)2– formed by interstitial N and O.

2.8. Optical Properties of TiO2, 5%NT/TiO2, and 5%NAB/TiO2 Catalysts

The light-harvesting capacities of the TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts were evaluated via Ultraviolet-visible (UV–vis) spectroscopy, as shown in Figure 8a. TiO2 showed a typical absorption edge at 420 nm (corresponding to the 3.05 eV bandgaps in Figure 8b), suggesting the lack of capacity for a visible light response in TiO2. On the contrary, N doping enhanced the visible light absorption of 5%NT/TiO2 and 5%NAB/TiO2, for which the light absorption edge was ~520 nm. Therefore, the incorporation of nitrogen into the TiO2 lattice led to the formation of new energy states, as the N 2p position was more negative than the O 2p state, leading to a decrease in the energy bandgap and a shift in the optical absorption towards the visible light region.

2.9. Photoelectrochemical Measurements of TiO2, 5%NT/TiO2, and 5%NAB/TiO2 Catalysts

The photoelectrochemical (PEC) measurement was conducted to study the photocurrent responses and charge transfer behavior (Figure 9a). TiO2 exhibited a lower photocurrent density under visible light irradiation due to its narrow light absorption range. However, 5%NT/TiO2 and 5%NAB/TiO2 showed an enhanced photocurrent density under visible light, indicating that nitrogen doping allowed the 5%NT/TiO2 and 5%NAB/TiO2 to convert visible light into photoexcited charges. In the results of electrochemical impedance spectroscopy (EIS) (Figure 9b), the 5%NT/TiO2 and 5%NAB/TiO2 catalysts showed lower semicircles compared with TiO2, indicating lower charge transfer resistance and better charge separation [30,31]. Among them, 5%NT/TiO2 had the lowest resistance and the strongest photocurrent because of the smaller lattice spacing of 5%NAB/TiO2, as shown in the HRTEM image. In general, a smaller lattice spacing means that the atoms or ions are closer to each other and the electrons move more easily in the lattice, resulting in lower resistance and stronger photogenerated carriers [32,33].
Mott–Schottky (MS) plots were used for estimating the band structure features of the catalysts. TiO2, 5%NT/TiO2 and 5%NT/TiO2 can be considered as an n-type semiconductors according to the positive slope of the MS curve. The estimated flat band potential for 5%NT/TiO2 was –0.6 V vs. an Ag/AgCl electrode, which can be converted to –0.36 V vs. a normal hydrogen electrode (NHE). In general, the conduction band (ECB) edge of typical n-type semiconductors is 0.20 V below its flat band potential. Then, the ECB value of 5%NT/TiO2 was determined to be −0.56 V vs. NHE. The valence band (EVB) potential of the 5%NT/TiO2 can be expressed by using the equation
E C B = E V B E g
In this way, the EVB position of 5%NT/TiO2 was calculated to be +2.54 V vs. NHE, while the EVB position of TiO2 and 5%NAB/TiO2 were calculated to be +3.09 and +2.86 V vs. NHE, respectively. The 5%NT/TiO2 showed the most negative ECB position compared with the potential of TiO2 (+0.09 V vs. NHE) and 5%NAB/TiO2 (–0.19 V vs. NHE). Moreover, the 5%NT/TiO2 showed a more negative ECB position (–0.56 V vs. NHE) compared with the potential of O2/·O2 (–0.33 V vs. NHE), suggesting that the photogenerated electrons of 5%NT/TiO2 could capture dissolved O2 to generate ·O2. Moreover, the EVB value of 5%NT/TiO2 was +2.54 V vs. NHE, which was higher than the potentials of OH/·OH (+2.40 V vs. NHE) and H2O/·OH (+2.38 V vs. NHE), indicating that photogenerated holes of 5%NT/TiO2 could directly yield ·OH radicals.

2.10. Photocatalytic MB Degradation Activity

It should be noted that the activity of 5%NAB/TiO2 was lower than that of 5%NT/TiO2 (Figure 10a), suggesting that interstitial doping was more effective than substitutive doping. This is because the interstitial doped N atoms were located in the lattice gaps of TiO2. XRD and HRTEM showed that this doping method reduced the lattice spacing of TiO2, meaning that the atoms were arranged more closely. This is beneficial for electron transfer, and this fact could be proved by the EIS curves. TEM and BET showed that interstitial doping made the grain size of TiO2 smaller and the specific surface area larger. In addition, the UV–vis and Mott–Schottky curves indicated that interstitial doped TiO2 had a stronger visible light absorption ability and adjusted its energy band structure to obtain stronger photogenerated carriers. This was confirmed by the photoinduced i–t curves.
At present, many studies have carried out various modifications of TiO2 for the degradation of methylene blue dye. For comparison, a summary of very recent studies on the photodegradation of MB by different TiO2-based photocatalysts is listed in Table 2. The percentages of degradation obtained by doped TiO2 prepared under our conditions were very near to those reported in the literature for TiO2 catalysts. Moreover, in our work, 5%NT/TiO2 still had good degradation performance for other organic dyes. As shown in Figure 10b, under the same reaction conditions, 5%NT/TiO2 could decolorize 59.0% of rhodamine B in 150 min or 50.0% of methyl orange. It was demonstrated that 5%NT/TiO2 has good general applicability. In addition, to explore the stability of the catalyst, we carried out SEM and XRD characterization analyses on the catalyst after 150 min of reaction with methylene blue. As shown in Figure 11a–e, the morphology, element distribution, and crystal form of the catalyst after the reaction were almost the same as those before the reaction. This indicated that the catalyst had good stability.
Previous studies demonstrated that the photocatalytic reaction pathway is believed to involve the reaction of MB with the generated ·OH radicals, producing a range of intermediate products to reach complete mineralization with the formation of CO2 and H2O [41]. To study the main active components in the degradation process of MB on the surface of TiO2 and to understand the degradation mechanism in more detail, trapping experiments were carried out. Silver nitrate, methanol, and isopropanol were used as scavengers to capture photogenerated electrons (e), photogenerated holes (h+), and hydroxyl radicals (·OH), respectively. To demonstrate the involvement of these radicals, a mixture of MB and silver nitrate (2% v/v) or methanol (2% v/v) or isopropanol (2% v/v) was irradiated under the same conditions. The results thus obtained are shown in Figure 12a. As can be observed, the addition of a photogenerated electron scavenger inhibited the degradation of MB. These results indicated that holes were the primary active species in the degradation of MB, while ·OH and photogenerated holes (h+) radicals were likely of secondary importance in photodegradation.
According to the results above and literature reports [42], the possible photocatalytic mechanism of the 5%NT/TiO2 photocatalyst was plotted as shown in Figure 12b. The photodegradation process depends on the generation and separation of carriers; under light conditions, once the semiconductor absorbs energy that is higher than the energy of its energy band, electrons will be excited from the valence band across the forbidden band to the conduction band. The doping of interstitial N reduced the width of the forbidden band of TiO2, which greatly increased the density of photogenerated carriers. O2 adsorbed on the surface of TiO2 formed active substances (·O2) after accepting electrons. At the same time, holes remained on the former, and the formed holes reacted directly with MB molecules, and H2O was adsorbed on the TiO2 surface to form the active species (·OH).

3. Materials and Methods

3.1. Preparation of the Catalysts

3.1.1. Preparation of 5%NT/TiO2

First, 40 mL of tetrabutyl titanate (Aladdin, Shanghai, China) was uniformly dispersed in 50 mL of an ethanol solution. It was vigorously stirred at 60 °C for 2 h and was named Solution A. Subsequently, 3 mL of nitric acid was added to 30 mL of distilled water and vigorously stirred for 5 min, and this was named Solution B. Immediately after that, thiourea (Aladdin, Shanghai, China) was added to 20 mL of ethanol and thoroughly stirred for 5 min. Subsequently, 3 mL of nitric acid was added and stirred at 60 °C for 1 h, and this was named Solution C. Then, under stirring, Solution B and Solution C were poured into Solution A at intervals of 5 min successively. Finally, the obtained mixture was poured into the polytetrafluoroethylene liner in the autoclave and maintained at 100 °C for 24 h. After centrifugal washing with ethanol, the sample was heated to 150 °C at a rate of 2 °C/min and maintained at this temperature for 1 h to remove the residual ethanol. After this stage was complete, the sample continued to be heated at the same rate until it reached 450 °C and was maintained at this temperature for 2 h.

3.1.2. Preparation TiO2 and 5%NAB/TiO2

The preparation methods of 5%NAB/TiO2 and TiO2 were the same as that for 5%NT/TiO2. The only difference was that the nitrogen precursor doped in 5%NAB/TiO2 was ammonium bicarbonate (Aladdin, Shanghai, China), and for TiO2, no nitrogen precursor was added.

3.2. Characterization of the Catalysts

The crystalline structure and phase composition of the TiO2, 5%NT/TiO2, and 5%NT/TiO2 catalysts were determined by X-ray diffraction (XRD) on an X-Pert diffractometer (Rigaku, Kyoto, Japan) equipped with graphite monochromatized Cu-Kα radiation. The elemental composition, chemical states, and surface chemistry of TiO2, 5%NT/TiO2, and 5%NT/TiO2 were analyzed using Agilent 5100 X-ray Photoelectron Spectroscopy (XPS, Agilent Technologies, Dallastown, PA, USA), where the tube’s voltage was 15 kV, the tube’s current was 10 mA, and an ultra-high vacuum chamber with a mu-metal magnetic shield was used. The specific surface areas were determined using a surface area analyzer (Micromeritics, Norcross, GA, USA) by the Brunauer–Emmett–Teller (BET) method. The morphology and size of the TiO2, 5%NT/TiO2, and 5%NT/TiO2 catalysts were observed using a transmission electron microscope (TEM-16-TS-008, Thermo Fisher Scientific, Waltham, MA, USA). The acceleration voltage was 200 kV, the point resolution was 0.248 nm, and the maximum magnification was 1.05 million times to visualize the morphology, size, and surface features of the TiO2, 5%NT/TiO2, and 5%NT/TiO2 catalysts. Fourier transform infrared spectroscopy (FTIR, Bruker Vertex 70 infrared spectrometer, Bruker Corporation, Billerica, MA, USA) was used to analyze the chemical bonds and functional groups present on the surface of the TiO2, 5%NT/TiO2 and 5%NT/TiO2 catalysts. FTIR spectra in transmission mode were recorded at a resolution of 4 cm-1 in the range of 400 to 4000 cm−1 using the KBr tabbing technique. The light absorption properties of the TiO2, 5%NT/TiO2 and 5%NT/TiO2 catalysts were determined by UV–vis spectroscopy with a wavelength range from 200 to 800 nm on a Japanese Shimadzu UV-3600i Plus Spectrometer. The photocurrent and electrochemical impedance spectroscopy were measured on an electrochemical workstation (Shanghai Chenhua Instrument Co., Ltd., Shanghai, China).

3.3. Evaluation of the Catalysts

The performance of different catalysts in the photocatalytic degradation of MB was evaluated in a photocatalytic reactor under visible light irradiation (300 W xenon lamp). Firstly, 0.02 g of each catalyst was added to 20 mL MB. Adsorption and desorption equilibrium were attained by agitating the suspensions with a magnetic stirrer for 1 h. Afterward, the condensate water was turned on to keep the reaction temperature at room temperature. Then the solution was irradiated with a 300 W xenon lamp without a light intensity meter as the source of visible light. An aliquot was collected periodically with a syringe from the reactor every 30 min during a 150 min interval. Finally, the catalyst was filtered and recycled. Changes in the concentration of MB were observed from its characteristic absorption at 664 nm using a UV–vis spectrometer. The degradation efficiency of MB at each time point was calculated using the following formula [43,44]
Degradation   efficiency   ( % ) = C o C t C o × 100 %
where Co is the initial concentration of MB, and Ct is the concentration of MB at a time.

4. Conclusions

In conclusion, this work provides a straightforward method to fabricate nitrogen-doped TiO2 photocatalysts based on thiourea as the N source. The systematic characterizations confirmed the presence of nitrogen in TiO2 crystals. UV–vis spectra showed the extended light absorption range on 5%NT/TiO2. Thus, 5%NT/TiO2 had a better photocurrent response during PEC measurement. Through optimizing the nitrogen concentration, 5%NT/TiO2 showed higher MB degradation activity than TiO2 and 5%NAB/TiO2 under visible light illumination. Our study clarified the advantages of interstitial N doping in improving photocatalytic activity under visible light conditions. More importantly, this work demonstrated that the introduction of nitrogen facilitates the utilization of solar light energy and makes TiO2 a promising material for environmental remediation and beyond.

Author Contributions

D.L.: software, data curation, and writing—review and editing. V.C.C.: validation and writing—original draft preparation. Y.L. (Yuqiao Li): validation and visualization. H.L.: supervision, writing—review, and funding. Y.L. (Yiming Lei): conceptualization, supervision, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work received financial support from the Young Talent Plan of Liaoning Province (XLYC2203068), the Scientific Research Foundation of Technology Department of Liaoning Province of China (2022-MS-379), and the National Natural Science Foundation of China (21902116).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The program of China Scholarships Council (No. 202206250016) is acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

X%NT/TiO2: x% represents the amount of N doping, and T indicates that the selected N source was thiourea. X% NAB/TiO2: x% represents the amount of N doping, and AB indicates that the selected N source was ammonium bicarbonate.

References

  1. Sadek, O.; Touhtouh, S.; Dahbi, A.; Hajjaji, A. Photocatalytic degradation of methylene blue on multilayer TiO2 coatings elaborated by the sol-gel spin-coating method. Water. Air. Soil. Pollut. 2023, 234, 698. [Google Scholar] [CrossRef]
  2. Ma, J.; Tian, Z.; Li, L.; Lu, Y.; Xu, X.; Hou, J. Loading nano-CuO on TiO2 nanomeshes towards efficient photodegradation of methylene blue. Catalysts 2022, 12, 383. [Google Scholar] [CrossRef]
  3. Ma, Y.; Tao, L.; Bai, S.; Hu, A. Green synthesis of Ag nanoparticles for plasmon-assisted photocatalytic degradation of methylene blue. Catalysts 2021, 11, 1499. [Google Scholar] [CrossRef]
  4. Hariharalakshmanan, R.K.; Watanabe, F.; Karabacak, T. In situ growth and UV photocatalytic effect of ZnO nanostructures on a Zn plate immersed in methylene blue. Catalysts 2022, 12, 1657. [Google Scholar] [CrossRef]
  5. Al-Ghouti, M.A.; Dib, S.S. Utilization of nano-olive stones in environmental remediation of methylene blue from water 03 chemical sciences 0306 physical chemistry (incl. structural). J. Environ. Heal. Sci. Eng. 2020, 18, 63–77. [Google Scholar] [CrossRef]
  6. Haq, F.; Kiran, M.; Chinnam, S.; Farid, A.; Khan, R.U.; Ullah, G.; Aljuwayid, A.M.; Habila, M.A.; Mubashir, M. Synthesis of bioinspired sorbent and their exploitation for methylene blue remediation. Chemosphere 2023, 321, 138000. [Google Scholar] [CrossRef]
  7. Xu, Y.; Shi, X.; Hua, R.; Zhang, R.; Yao, Y.; Zhao, B.; Liu, T.; Zheng, J.; Lu, G. Remarkably catalytic activity in reduction of 4-nitrophenol and methylene blue by Fe3O4@COF supported noble metal nanoparticles. Appl. Catal. B Environ. 2020, 260, 118142. [Google Scholar] [CrossRef]
  8. Malato, S.; Fernández-Ibáñez, P.; Maldonado, M.I.; Blanco, J.; Gernjak, W. Decontamination and disinfection of water by solar photocatalysis: Recent overview and trends. Catal. Today 2009, 147, 1–59. [Google Scholar] [CrossRef]
  9. Bairamis, F.; Konstantinou, I.; Petrakis, D.; Vaimakis, T. Enhanced performance of electrospun nanofibrous TiO2/g-C3N4 photocatalyst in photocatalytic degradation of methylene blue. Catalysts 2019, 9, 880. [Google Scholar] [CrossRef]
  10. Babyszko, A.; Wanag, A.; Sadłowski, M.; Kusiak-Nejman, E.; Morawski, A.W. Synthesis and characterization of SiO2/TiO2 as photocatalyst on methylene blue degradation. Catalysts 2022, 12, 1372. [Google Scholar] [CrossRef]
  11. Zhang, C.; Zhou, Y.; Bao, J.; Sheng, X.; Fang, J.; Zhao, S.; Zhang, Y.; Chen, W. Hierarchical honeycomb Br-, N-codoped TiO2 with enhanced visible-light photocatalytic H2 production. ACS Appl. Mater. Interfaces 2018, 10, 18796–18804. [Google Scholar] [CrossRef] [PubMed]
  12. Assayehegn, E.; Solaiappan, A.; Chebude, Y.; Alemayehu, E. Fabrication of tunable anatase/rutile heterojunction N/TiO2 nanophotocatalyst for enhanced visible light degradation activity. Appl. Surf. Sci. 2020, 515, 145966. [Google Scholar] [CrossRef]
  13. Chakraborty, A.K.; Ganguli, S.; Sabur, M.A. Nitrogen doped titanium dioxide (N-TiO2): Electronic band structure, visible light harvesting and photocatalytic applications. J. Water Process Eng. 2023, 55, 104183. [Google Scholar] [CrossRef]
  14. Bissinger, D.; Honerkamp, J.H.; Roldan, J.; Bremes, J.; Kannen, K.; Lake, M.K.; Roppertz, A. Development of catalytically functionalized polyester-based filters produced by flame spray pyrolysis. Top. Catal. 2024, 67, 539–550. [Google Scholar] [CrossRef]
  15. Khan, T.T.; Rafiqul Bari, G.A.K.M.; Kang, H.J.; Lee, T.G.; Park, J.W.; Hwang, H.J.; Hossain, S.M.; Mun, J.S.; Suzuki, N.; Fujishima, A.; et al. Synthesis of N-doped TiO2 for efficient photocatalytic degradation of atmospheric NOx. Catalysts 2021, 11, 109. [Google Scholar] [CrossRef]
  16. Das, D.; Shyam, S. Reduced work function in anatase ⟨101⟩ TiO2 films self-doped by O-vacancy-dependent Ti3+ bonds controlling the photocatalytic dye degradation performance. Langmuir 2024, 40, 10502–10517. [Google Scholar] [CrossRef]
  17. Divyasri, Y.V.; Lakshmana Reddy, N.; Lee, K.; Sakar, M.; Navakoteswara Rao, V.; Venkatramu, V.; Shankar, M.V.; Gangi Reddy, N.C. Optimization of N doping in TiO2 nanotubes for the enhanced solar light mediated photocatalytic H2 production and dye degradation. Environ. Pollut. 2020, 269, 116170. [Google Scholar] [CrossRef]
  18. Bhowmick, S.; Saini, C.P.; Santra, B.; Walczak, L.; Semisalova, A.; Gupta, M.; Kanjilal, A. Modulation of the work function of TiO2 nanotubes by nitrogen doping: Implications for the photocatalytic degradation of dyes. ACS Appl. Nano Mater. 2023, 6, 50–60. [Google Scholar] [CrossRef]
  19. Zhang, X.; Zuo, G.; Lu, X.; Tang, C.; Cao, S.; Yu, M. Anatase TiO2 sheet-assisted synthesis of Ti3+ self-doped mixed phase TiO2 sheet with superior visible-light photocatalytic performance: Roles of anatase TiO2 sheet. J. Colloid. Interface Sci. 2017, 490, 774–782. [Google Scholar] [CrossRef]
  20. Sathish, M.; Viswanathan, B.; Viswanath, R.P. Alternate synthetic strategy for the preparation of CdS nanoparticles and its exploitation for water splitting. Int. J. Hydrog. Energy 2006, 31, 891–898. [Google Scholar] [CrossRef]
  21. Chen, X.; Burda, C. Photoelectron spectroscopic investigation of nitrogen-doped titania nanoparticles. J. Phys. Chem. B 2004, 108, 15446–15449. [Google Scholar] [CrossRef]
  22. Pustovalova, A.A.; Pichugin, V.F.; Ivanova, N.M.; Bruns, M. Structural features of N-containing titanium dioxide thin films deposited by magnetron sputtering. Thin Solid. Film. 2017, 627, 9–16. [Google Scholar] [CrossRef]
  23. Ren, W.; Ai, Z.; Jia, F.; Zhang, L.; Fan, X.; Zou, Z. Low temperature preparation and visible light photocatalytic activity of mesoporous Carbon-Doped Crystalline TiO2. Appl. Catal. B Environ. Energy 2007, 69, 138–144. [Google Scholar] [CrossRef]
  24. Di Valentin, C.; Pacchioni, G.; Selloni, A.; Livraghi, S.; Giamello, E. Characterization of paramagnetic species in N-doped TiO2 powders by EPR spectroscopy and DFT calculations. J. Phys. Chem. B 2005, 109, 11414–11419. [Google Scholar] [CrossRef]
  25. Li, B.; Dai, F.; Xiao, Q.; Yang, L.; Shen, J.; Zhang, C.; Cai, M. Nitrogen-doped activated carbon for a high energy hybrid supercapacitor. Energy Environ. Sci. 2015, 9, 102–106. [Google Scholar] [CrossRef]
  26. Zhang, Z.; Lu, S.; Shen, G.; Zhao, Y.; Zhu, T.; Gao, Q.; Sun, N.; Wei, W. Controllable and rapid synthesis of nitrogen-doped ordered mesoporous carbon single crystals for CO2 capture. J. CO2 Util. 2021, 56, 101851. [Google Scholar] [CrossRef]
  27. Rustamaji, H.; Prakoso, T.; Devianto, H.; Widiatmoko, P.; Saputera, W.H. Urea nitrogenated mesoporous activated carbon derived from oil palm empty fruit bunch for high-performance supercapacitor. J. Energy Storage 2022, 52, 104724. [Google Scholar] [CrossRef]
  28. Wang, Y.; Yin, C.; Qin, H.; Wang, Y.; Li, Y.; Li, X.; Zuo, Y.; Kang, S.; Cui, L. A urea-assisted template method to synthesize mesoporous N-doped CeO2 for CO2 capturet. Dalt. Trans. 2015, 44, 18718–18722. [Google Scholar] [CrossRef]
  29. Ghosh, A.; Ghosh, S.; Seshadhri, G.M.; Ramaprabhu, S. Green synthesis of nitrogen-doped self-assembled porous carbon-metal oxide composite towards energy and environmental applications. Sci. Rep. 2019, 9, 5187. [Google Scholar] [CrossRef]
  30. Jiang, Z.; Ran, M.; Liu, K.; Huang, Y.; Li, Z.; Shen, T.; Li, W.; Khojiev, S.; Hu, Z.; Liu, J. Directing Ni/Al layered double hydroxides nanosheets on tubular graphite carbon nitride for promoted photocatalytic hydrogen production. Mater. Today Chem. 2024, 39, 102135. [Google Scholar] [CrossRef]
  31. Jiang, L.; Yu, H.; Shi, L.; Zhao, Y.; Wang, Z.; Zhang, M.; Yuan, S. Optical band structure and photogenerated carriers transfer dynamics in FTO/TiO2 heterojunction photocatalysts. Appl. Catal. B Environ. 2016, 199, 224–229. [Google Scholar] [CrossRef]
  32. Li, Y.; Lai, R.; Luo, X.; Liu, X.; Ding, T.; Lu, X.; Wu, K. On the absence of a phonon bottleneck in strongly confined CsPbBr 3 perovskite nanocrystals. Chem. Sci. 2019, 10, 5983–5989. [Google Scholar] [CrossRef]
  33. Zhu, Y.; Cui, Q.; Chen, J.; Chen, F.; Shi, Z.; Zhao, X.; Xu, C. Inhomogeneous trap-state-mediated ultrafast photocarrier dynamics in CsPbBr3 microplates. ACS Appl. Mater. Interfaces 2021, 13, 6820–6829. [Google Scholar] [CrossRef]
  34. Kumar, M.R.A.; Abebe, B.; Nagaswarupa, H.P.; Murthy, H.C.A.; Ravikumar, C.R.; Sabir, F.K. Enhanced photocatalytic and electrochemical performance of TiO2-Fe2O3 nanocomposite: Its applications in dye decolorization and as supercapacitors. Sci. Rep. 2020, 10, 1249. [Google Scholar] [CrossRef]
  35. Wan, J.; Wei, M.; Hu, Z.; Peng, Z.; Wang, B.; Feng, D.; Shen, Y. Ternary composites of TiO2 nanotubes with reduced graphene oxide (RGO) and meso-tetra (4-carboxyphenyl) porphyrin for enhanced visible light photocatalysis. Int. J. Hydrog. Energy 2016, 33, 14692–14703. [Google Scholar] [CrossRef]
  36. Ling, M.F.C.; Hui, K.C.; Sambudi, N.S. Modification of TiO2 with clam-shell powder for photodegradation of methylene blue. J. Sol. Gel Sci. Technol. 2022, 102, 412–421. [Google Scholar] [CrossRef]
  37. Kim, B.C.; Jeong, E.; Kim, E.; Hong, S.W. Bio-organic-inorganic hybrid photocatalyst, TiO2 and glucose oxidase composite for enhancing antibacterial performance in aqueous environments. Appl. Catal. B Environ. Energy 2018, 242, 194–201. [Google Scholar] [CrossRef]
  38. Magnone, E.; Kim, M.-K.; Lee, H.J.; Park, J.H. Testing and substantial improvement of TiO2/UV photocatalysts in the degradation of methylene blue. Ceram. Int. 2019, 10, 3359–3367. [Google Scholar] [CrossRef]
  39. Ridha, N.J.; Alosfur, F.K.M.; Kadhim, H.B.A.; Ahmed, L.M. Synthesis of Ag decorated TiO2 nanoneedles for photocatalytic degradation of methylene blue dye. Mater. Res. Express 2021, 8, 125013. [Google Scholar] [CrossRef]
  40. Abbas, F.; Bensaha, R. Effect of annealing time on structural and optical proprieties of mercury (Hg+2) doped TiO2 thin films elaborated by sol-gel method for future photo-catalytic application. Opt. 2021, 247, 167846. [Google Scholar] [CrossRef]
  41. da Silva, C.G.; Faria, J.L. Photochemical and photocatalytic degradation of an azo dye in aqueous solution by UV irradiation. J. Photochem. Photobiol. A Chem. 2003, 155, 133–143. [Google Scholar] [CrossRef]
  42. Zhou, Q.; Zhang, L.; Zuo, P.; Wang, Y.; Yu, Z. Enhanced photocatalytic performance of spherical BiOI/MnO2 composite and mechanism investigation. RSC Adv. 2018, 8, 36161–36166. [Google Scholar] [CrossRef] [PubMed]
  43. Trandafilović, L.V.; Jovanović, D.J.; Zhang, X.; Ptasińska, S.; Dramićanin, M.D. Enhanced photocatalytic degradation of methylene blue and methyl orange by ZnO: Eu nanoparticles. Appl. Catal. B Environ. 2017, 203, 740–752. [Google Scholar] [CrossRef]
  44. Girotto, G.Z.; Thill, A.S.; Matte, L.P.; Vogt, M.A.H.; Machado, T.V.; Dick, L.F.P.; Mesquita, F.; Bernardi, F. Ni/SrTiO3 nanoparticles for photodegradation of methylene blue. ACS Appl. Nano Mater. 2022, 5, 13295–13307. [Google Scholar] [CrossRef]
Figure 1. The degradation rates of an MB solution by NT/TiO2 photocatalysts with different amounts of N doping.
Figure 1. The degradation rates of an MB solution by NT/TiO2 photocatalysts with different amounts of N doping.
Catalysts 14 00681 g001
Figure 2. XRD patterns of (a) TiO2, 5%NT/TiO2, and 5%NAB/TiO2, and (b) local magnification image of XRD.
Figure 2. XRD patterns of (a) TiO2, 5%NT/TiO2, and 5%NAB/TiO2, and (b) local magnification image of XRD.
Catalysts 14 00681 g002
Figure 3. (a) SEM image and (b) the corresponding Ti element, (c) O element and (d) N element mappings of 5%NT/TiO2.
Figure 3. (a) SEM image and (b) the corresponding Ti element, (c) O element and (d) N element mappings of 5%NT/TiO2.
Catalysts 14 00681 g003
Figure 4. XPS survey spectra. (a) Ti 2p XPS spectra, (b) O 1s spectra, and (c) N 1s spectra of the catalysts.
Figure 4. XPS survey spectra. (a) Ti 2p XPS spectra, (b) O 1s spectra, and (c) N 1s spectra of the catalysts.
Catalysts 14 00681 g004
Figure 5. (a,d) TEM and HRTEM images for TiO2, (b,e) 5%NT/TiO2, and (c,f) 5%NAB/TiO2. R, rutile; A, anatase.
Figure 5. (a,d) TEM and HRTEM images for TiO2, (b,e) 5%NT/TiO2, and (c,f) 5%NAB/TiO2. R, rutile; A, anatase.
Catalysts 14 00681 g005
Figure 6. (a) N2 adsorption–desorption isotherms and (b) pore size distribution of the catalysts.
Figure 6. (a) N2 adsorption–desorption isotherms and (b) pore size distribution of the catalysts.
Catalysts 14 00681 g006
Figure 7. FTIR spectra of pure TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts.
Figure 7. FTIR spectra of pure TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts.
Catalysts 14 00681 g007
Figure 8. (a) UV–vis spectra of the TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts. (b) Bandgaps of the TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts.
Figure 8. (a) UV–vis spectra of the TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts. (b) Bandgaps of the TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts.
Catalysts 14 00681 g008
Figure 9. (a) Photoinduced i–t curves, (b) EIS curves, and (c) Mott–Schottky plots for the TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts.
Figure 9. (a) Photoinduced i–t curves, (b) EIS curves, and (c) Mott–Schottky plots for the TiO2, 5%NT/TiO2, and 5%NAB/TiO2 catalysts.
Catalysts 14 00681 g009
Figure 10. (a) Comparison of the rates of MB degradation by TiO2, 5%NT/TiO2, and 5%NAB/TiO2 photocatalysts. (b) The degradation rates of rhodamine B and methyl orange by 5%NT/TiO2 photocatalysts.
Figure 10. (a) Comparison of the rates of MB degradation by TiO2, 5%NT/TiO2, and 5%NAB/TiO2 photocatalysts. (b) The degradation rates of rhodamine B and methyl orange by 5%NT/TiO2 photocatalysts.
Catalysts 14 00681 g010
Figure 11. (a) XRD patterns of spent 5%NT/TiO2. (b) SEM image and (c) the corresponding Ti element, (d) O element, and (e) N element mappings of spent 5%NT/TiO2.
Figure 11. (a) XRD patterns of spent 5%NT/TiO2. (b) SEM image and (c) the corresponding Ti element, (d) O element, and (e) N element mappings of spent 5%NT/TiO2.
Catalysts 14 00681 g011
Figure 12. (a) Effects of the addition of radical scavengers on the photodegradation of MB and (b) schematic diagram of the photocatalytic mechanism of the NT/TiO2 photocatalyst.
Figure 12. (a) Effects of the addition of radical scavengers on the photodegradation of MB and (b) schematic diagram of the photocatalytic mechanism of the NT/TiO2 photocatalyst.
Catalysts 14 00681 g012
Table 1. BET surface area and average pore size of the catalysts.
Table 1. BET surface area and average pore size of the catalysts.
CatalystsSurface Area (m2·g−1)Average Pore Size (nm)
TiO237.09.5
5%NT/TiO2107.73.5
5%NAB/TiO290.09.6
Table 2. Photocatalytic activity of recently studied TiO2-based photocatalysts for MB degradation.
Table 2. Photocatalytic activity of recently studied TiO2-based photocatalysts for MB degradation.
Irradiation Time (min)Degradation Efficiency (%)Ref.
5%NT/TiO215056.5This work
TiO2–Fe2O36062.9[34]
P2512020.0[35]
TiO2/CS18033.9[36]
TiO2-GOx60 50.0[37]
AlHF-TiO26030.0[38]
Ag-TiO25040.0[39]
Hg-doped TiO212056.7[40]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, D.; Calebe, V.C.; Li, Y.; Liu, H.; Lei, Y. Interstitial N-Doped TiO2 for Photocatalytic Methylene Blue Degradation under Visible Light Irradiation. Catalysts 2024, 14, 681. https://doi.org/10.3390/catal14100681

AMA Style

Li D, Calebe VC, Li Y, Liu H, Lei Y. Interstitial N-Doped TiO2 for Photocatalytic Methylene Blue Degradation under Visible Light Irradiation. Catalysts. 2024; 14(10):681. https://doi.org/10.3390/catal14100681

Chicago/Turabian Style

Li, Dezheng, Vilanculo Clesio Calebe, Yuqiao Li, Huimin Liu, and Yiming Lei. 2024. "Interstitial N-Doped TiO2 for Photocatalytic Methylene Blue Degradation under Visible Light Irradiation" Catalysts 14, no. 10: 681. https://doi.org/10.3390/catal14100681

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