Next Article in Journal
Fabrication of UV-Curable Polysiloxane Coating with Tunable Refractive Index Based on Controllable Hydrolysis
Next Article in Special Issue
Construction of Built-In Electric Field in TiO2@Ti2O3 Core-Shell Heterojunctions toward Optimized Photocatalytic Performance
Previous Article in Journal
Photoluminescence of Two-Dimensional MoS2 Nanosheets Produced by Liquid Exfoliation
Previous Article in Special Issue
A Review on Nano Ti-Based Oxides for Dark and Photocatalysis: From Photoinduced Processes to Bioimplant Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of Z-Type TiN@(A,R)TiO2 Plasmonic Photocatalyst with Enhanced Photocatalytic Activity

1
Shaanxi Key Laboratory of Green Preparation and Functionalization for Inorganic Materials, School of Material Science and Engineering, Shaanxi University of Science & Technology, Xi’an 710021, China
2
State Key Laboratory for Mechanical Behavior of Materials, School of Materials Science and Engineering, Xi’an Jiaotong University, Xi’an 710049, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(13), 1984; https://doi.org/10.3390/nano13131984
Submission received: 15 May 2023 / Revised: 26 June 2023 / Accepted: 27 June 2023 / Published: 30 June 2023
(This article belongs to the Special Issue Synthesis of TiO2 Nanoparticles and Their Catalytic Activity)

Abstract

:
Plasmonic effect-enhanced Z-type heterojunction photocatalysts comprise a promising solution to the two fundamental problems of current TiO2-based photocatalysis concerning low-charge carrier separation efficiency and low utilization of solar illumination. A plasmonic effect-enhanced TiN@anatase-TiO2/rutile-TiO2 Z-type heterojunction photocatalyst with the strong interface of the N–O chemical bond was synthesized by hydrothermal oxidation of TiN. The prepared photocatalyst shows desirable visible light absorption and good visible-light-photocatalytic activity. The enhancement in photocatalytic activities contribute to the plasma resonance effect of TiN, the N–O bond-connected charge transfer channel at the TiO2/TiN heterointerface, and the synergistically Z-type charge transfer pathway between the anatase TiO2 (A-TiO2) and rutile TiO2 (R-TiO2). The optimization study shows that the catalyst with a weight ratio of A-TiO2/R-TiO2/TiN of approximately 15:1:1 achieved the best visible light photodegradation activity. This work demonstrates the effectiveness of fabricating plasmonic effect-enhanced Z-type heterostructure semiconductor photocatalysts with enhanced visible-light-photocatalytic activities.

1. Introduction

Green nanotechnology driven by solar energy has attracted great interest in alleviating the environmental hazards of pesticides, organic dyes, toxic gases, and industrial wastewater [1,2,3,4,5,6]. Since Carey et al. [7,8] used semiconductors to degrade pollutants in 1976, TiO2 has proven to be a material that can be used for environmental purification. However, pure TiO2 possesses a wide band gap (about 3.2 eV) [9,10]. Due to this limitation, it only responds to UV light and has low solar energy utilization (about 4%). Furthermore, the recombination rate of photogenerated charge carriers generated after TiO2 excitation is much higher than that of interfacial charge transfer, resulting in low activity even under UV light [11]. Therefore, promoting solar utilization and charge carrier separation is the key to improving the photocatalytic performance of the catalysts.
Combining the localized surface plasmon resonance (LSPR) effect with semiconductor photocatalysts is a promising method to promote both the charge carrier separation efficiency and the responsive solar illumination range [12]. Till now, most plasmonic photocatalysts relied on noble metal nanostructures (such as Au or Ag) [13]. However, their potential for practical applications is limited due to their rarity, high cost, low thermal stability, and easy dissolution upon the exposure to air or humidity. Thus, novel plasmonic photocatalysts without noble metal components should be developed to overcome these problems.
Recently, TiN has emerged as an attractive competitor in photocatalytic applications due to its plasmonic resonance absorption properties [14]. In addition, the work function of TiN is about 4 eV versus vacuum [15], which is greater than or equal to the electron affinity of most semiconductor metal oxide photocatalysts, including TiO2. Therefore, TiN tends to form a favorable energetic alignment to promote hot carrier-enhanced solar energy conversion [16]. Naldoni et al. [17] explored the plasmonic-enhanced TiO2 photocatalysts by coupling with TiN, demonstrating that the LSPR effect of the TiN introduced an enhanced photocurrent generation and photocatalytic activity. Fakhouri et al. [18] demonstrated a significant photoactive improvement to bilayered RF magnetron-sputtering TiN/TiO2 thin films due to enhanced charge separation at the heterojunction. Clatworthy et al. [19] demonstrated enhanced photocatalytic activity of TiN-TiO2 nanoparticle composites and proposed that hot electrons migration can be promoted due to TiO2 photovoltage by combining visible light with UV light. However, there is usually a certain lattice mismatch between different semiconductors, therefore constructed heterostructures usually result in large lattice defects and interface resistance [20]. These lattice defects often form the capturing center of photogenerated carriers [21], and the interface resistance would restrict charge transfer and affect their stability [22], thus greatly affecting the efficiency of charge carrier separation. Therefore, a novel plasmonic photocatalyst without noble metal components could be developed if a nanostructured TiN/TiO2 composite with good contact could be created. Zhu et al. [23] found that the epitaxial growth of different semiconductors on conductive precursors and the regulation of growth conditions can significantly reduce the interface contact resistance, which can solve the challenge of building heterostructures to obtain high photogenerated charge separation characteristics. Li et al. [24] fabricated a TiN/TiO2 plasmonic photocatalyst by in situ growing TiO2 on TiN nanoparticles, demonstrating good visible light photocatalytic performance. Furthermore, Zhang et al. [25] significantly reduced the interface resistance and greatly improved their ability to photoelectrochemical decompose water by forming a strong interface contact of the S–O covalent bond at the interface of the Cu2S/Fe2O3 heterostructure. In our previous work [26], the Ti–O–Zr bonded TiO2/ZrTiO4 heterointerface was constructed by growing ZrTiO4 in situ on TiO2 to enhance the transport of photogenerated carriers. However, the possibility of forming a chemically bonded TiO2/TiN heterostructure and its synergistic enhancement of photocatalytic activities with the LSPR effect of TiN have yet to be explored.
In addition, fabrication of the direct Z-type heterojunction is an alternative strategy to obtain a semiconductor photocatalyst with high performance due to its advantage in charge carrier separation and utilizing the high-redox properties of each component [27]. For direct Z-type heterojunctions with staggered band structure, the photogenerated electrons on lower CB and the holes on the higher VB recombine. Meanwhile, the electrons and holes with stronger redox abilities are retained [28,29]. Thus, the charge carrier separation can be enhanced, and the highest redox potential of the heterojunction can be retained, thus contributing to the promoted photocatalytic activities. In our previous work, we successfully constructed a direct Z-type A-TiO2/R-TiO2 heterojunction by synergistically mediating oxygen vacancy contents and the band structure of the catalysts through B-doping [30]. However, to our best knowledge, there is no report concerning the possibility of combining TiN plasmonic enhancement with direct Z-type TiO2-based heterojunction.
In this study, to solve the two fundamental problems of current TiO2-based photocatalysis on low charge carrier separation efficiency and low utilization of solar illumination, we optimized a unique plasmonic effect-enhanced Z-type TiN@A-TiO2/R-TiO2 photocatalyst with a strong interface of the N–O chemical bond through hydrothermal in situ oxidation of TiN to (A,R)-TiO2. In this photocatalyst system, the desirable visible light absorption could be attributed to the LSPR effect of the TiN component. The charge carrier separation efficiency could be enhanced by the Z-type charge transfer mode at the interface of the A-TiO2/R-TiO2 heterojunction. The obtained TiN@A-TiO2/R-TiO2 photocatalyst showed a distinct enhancement in visible light absorption, photocurrent generation, and photodegradation activities, demonstrating a simple way to promote the photoactive properties of semiconductor photocatalysts by fabricating plasmonic effect-enhanced Z-type heterostructures.

2. Experimental Section

2.1. Chemicals

Commercial titanium nitride (TiN, AR) were procured from Aladdin Reagent Co., Ltd., Shanghai, China. Ethanol (C2H5OH, CP); hydrogen peroxide (H2O2, 35 wt%, AR), rhodamine B (RhB, AR), and concentrated sulfuric acid (H2SO4, AR) were procured from the National Reagent company, Beijing, China. All reagents were used as received.

2.2. Preparation of the Catalyst

All samples were prepared by a simple hydrothermal process. Firstly, TiN powder was dispersed in 40 mL of deionized water and sonicated for 15 min to obtain TiN suspension. A total of 1 mL of H2SO4 (1 M) and a certain amount of H2O2 was added dropwise and stirred for 2 h. The suspension was then hydrothermally treated at 180 °C for 5 h, then washed and dried at 60 °C for 24 h to obtain the target samples. The hydrolysis degree of TiN was determined by the amount of H2O2 added. In this work, the mass fraction of added H2O2 is 0%, 0.5%, 1.0%, 2.5%, and 5.0%. The obtained catalysts were labeled as TiN, sample 1 (S1), sample 2 (S2), sample 3 (S3), and sample 4 (S4), respectively.

2.3. Characterization Methods

Compositions were recorded on a D/max-2200PC powder X-ray diffraction (XRD), with Cu Kα radiation over a 2θ ranging from 10° to 70°. Morphologies and microstructures were recorded by SEM (FEI Verios 460, Hillsboro, OR, USA), TEM, and HRTEM (FEI Tecnai G2 F20 S-TWIN, Hillsboro, OR, USA). XPS was studied on an X-ray photoelectron spectroscope (XPS, AXIS SUPRA, Manchester, UK) with a monochromatic Al Kα source. Comparing with the standard binding energy of adsorbed carbon (284.6 eV), charge correction was applied after the peak fitting using the CasaXPS analysis software. UV–Vis diffuse reflectance spectra (DRS) were tested on a UV–Vis-NIR spectrophotometer (Cary 5000, Santa Clara, CA, USA). Photoluminescence (PL) spectra were obtained by a fluorescence spectrophotometer (F-4600, Rigaku, Japan) with the excitation source at 345 nm. EPR spectra were conducted by a Bruker A300 spectrometer, during which DMPO was applied using a 300 W Xe lamp as the light source. The FT-IR spectrum (4000–500 cm−1) was obtained on Vertex70 Bruker FT-IR Spectroscopy.
Photoelectrochemical analysis was performed on a CHI760D electrochemical workstation equipped with a 300 W Xe lamp and a cut off filter (>420 nm), in which 20 μL of catalyst slurry on an FTO substrate of 2 cm × 2 cm was used as the working electrode. The slurry was prepared by dispersing 5 mg catalyst powder into polyvinylidene difluoride N-methyl pyrrolidone solution (0.5 g, 2 wt%) through ultrasonic vibration. A total of 0.5 M Na2SO4 solution, platinum, and Ag/AgCl were used as the electrolyte solution, counter, and reference electrode, respectively.

2.4. Photocatalytic Performance

The photocatalytic activity of various catalysts was evaluated by RhB photodegradation. Firstly, 30 mL of RhB solution (10 mg/L) was prepared, then 30 mg of catalyst was added under stirring. Then, the above suspension was illuminated under visible light for 120 min, collected, centrifuged, and measured at regular intervals of 30 min. The peak absorbency of the centrifuged RhB solution at 554 nm was applied to analyze its concentration using a UV–Vis spectrophotometer.

3. Results and Discussion

3.1. Structural Characterization of the Photocatalysts

Figure 1 shows the XRD results of samples obtained with various H2O2 content. Except for the single-phase TiN sample, all other samples show diffraction peaks ascribed to three phases—that is, A-TiO2, R-TiO2, and TiN [31]. Moreover, as the H2O2 content increases, the peak intensity of TiN decreases and that of TiO2 increases. The content of each phase was determined by the Rietveld method and shown in Table 1. With an increase in the H2O2 content, the phase proportion of TiN gradually decreases accompanied by an increase in the TiO2 content, indicating the hydrolysis of TiN and its conversion into TiO2. Furthermore, the ratio of A-TiO2 to R-TiO2 increases accordingly. Specifically, for sample S2, the weight ratio of the three phases (A-TiO2:R-TiO2:TiN) is about 15:1:1.
Figure 2 shows the FT-IR spectra of sample TiN and S2. In the spectra, the wide absorption band at 3440 cm−1 and the peaks around 1633 cm−1 are ascribed to the adsorbed water and hydroxyl groups [32], respectively. The NOx-determined peaks appeared at 1382 cm−1 and 1346 cm−1 [33]. Furthermore, peaks between 500–800 cm−1 are believed to be caused by the stretching vibration of Ti–O–Ti bonds [34]. Compared to the TiN sample, the maximum strength of the Ti–O–Ti bond increased significantly, and a new peak of NOx appeared, indicating the formation of TiO2 and the possible existence of a newly formed N–O bond in sample S2.
Figures S1–S3 show the SEM images, particle size distributions, and BET surface areas of all the samples. As can be seen, all the samples show uniform and fine particle distribution with an average particle size of about 50 nm, except for sample TiN and S2, which show a slightly smaller size (around 35 nm). Moreover, except for the comparison samples (P25 and TiN), all the other samples have close BET surface areas, suggesting that surface area is not the reason for the photocatalytic performance difference between various samples. Microstructure was studied through TEM analysis. In the TEM results (Figure 3a), uniformly distributed irregular nanoparticles including polygonal, spherical, and rod-shaped particles can be observed. The length of rod particles is 50–200 nm, while the particle size of polygon and spherical particles is about 25 nm. In the HRTEM results (Figure 3b), spacings of 0.210, 0.324, and 0.352 nm of the lattice fringes, correspond to the (200) plane of TiN, the (110) plane of R-TiO2, and the (101) plane of A-TO2, respectively [35,36]. Moreover, the (A,R)-TiO2 are identified on the surface of TiN, and all three phases are in close contact. Furthermore, the energy spectra (Figure 3c–f) show the evenly distributed Ti, N, and O elements, indicating the potential formation of the TiN/(A,R)-TiO2 heterointerface. Therefore, it can be deduced that the in situ oxidation of TiN and growth of (A,R)-TiO2 could create heterojunctions with an intimate contact interface, improving charge transfer efficiency.
Figure 4a is the XPS survey spectrum of S2, demonstrating the presence of C, O, and Ti elements. The N element was not detected, suggesting that it may appear in the interior of the particles. The Ti 2p XPS spectrum was fitted into four peaks. The Ti 2p3/2 (458.1 eV) and the Ti 2p1/2 (464.1 eV) were for TiO2 [37,38,39,40]. The Ti 2p3/2 (457.3 eV) and Ti 2p1/2 (462.5 eV) were for partially oxidized TiN [41]. This observation confirms the creation of TiO2 from the oxidation of TiN and suggests the possibility of forming a chemical contact interface between TiN and TiO2. By further analysis of the O 1s spectrum of S2 (Figure 4c), four peaks can be fitted at 533.09, 531.54, 529.71, and 529.15 eV, which could be attributed to adsorbed H2O (AO), Ti–O in Ti2O3 suggests the existence of oxygen vacancies (VO), and oxygen in Ti–O–N and Ti–O lattices [42,43,44], respectively. Considering the relatively high area ratio of the Vo XPS peak, it can be deduced that there is a high content of VO in the sample. Based on the analysis of the O1s spectrum, it is clear that N–O bonds exist between TiN and its oxidation products TiO2, which contribute to the good contact interface of the formed TiN/TiO2 heterostructure.

3.2. Photodegradation Performance

Figure 5 shows the photodegradation performance of the catalysts. As can be seen, all samples showed no obvious adsorption in the dark reaction. The sample TiN did not have a degrading effect on RhB, indicating that it is not the main catalytic carrier in the photocatalytic reaction process but a cocatalyst. Compared to TiN samples, the hydrolyzed samples showed obvious degrading behavior on RhB. With the deeper degree of hydrolysis, i.e., the decrease in TiN content and the increase in weight ratio of A-TiO2 to R-TiO2, the photodegradation rate increases first and then decreases. Among them, the S2 sample has the best degradation efficiency, reaching more than 97% in 90 min. As for the kinetics of RhB degradation, the degradation curves are well-fitted by a mono-exponential curve, indicating that the photodegradation experiments follow the first-order kinetics [35]. Figure 5b shows the relationship between ln (C0/C) and t for all experiments using different samples, where C0 is the initial RhB content and C is the RhB concentration at reaction time t. By regression analysis of the linear curve in the graph, the value of the apparent first-order rate constant can be directly obtained, in which the value of sample S2 is the highest 0.02272 min−1. In addition, the cycling experiment (Figure 5c) shows that the sample can maintain a degradation efficiency of more than 90% after five cycles, showing good stability. From the XRD analysis in Figure 5d, no detectable differences can be seen between the as-prepared and cycled S2, indicating a well-preserved crystalline structure of the catalyst after multiple photocatalytic cycles. Moreover, a few studies on the RhB photodegradation performance of TiO2-based photocatalysts are summarized in Table 2. The table shows that the visible-light degradation performance of RhB over the catalyst prepared in this work was enhanced, indicating that the prepared TiN@(A,R)TiO2 is a promising visible-light photocatalyst.
Figure S4 shows the result of the free radical capture experiment. By adding IPA, TEOA, BQ, and AgNO3 as the capture agents of ·OH, h+, ·O2, and e, respectively, the effect of free radicals on photocatalysis was investigated [45]. The addition of TEOA and BQ has the greatest impact on the photodegradation rate, suggesting that the corresponding h+ and ·O2 may play the main role in the photodegradation process. EPR test was carried out on sample S2, and the result is shown in Figure 6. In the O2 free radical detection, the DMPO-·O2 signal peak of 1:1:1:1 was detected, and its intensity increased with prolonged irradiation time, confirming that the O2 radical is the main active species, whereas for ·OH radicals, no DMPO-·OH signal peak of 1:2:2:1 can be detected, suggesting that no ·OH radical can be produced during light irradiation. This result indirectly verified that h+ might participate in the following photodegradation reaction without conversion into ·OH.
Table 2. Summary of recent relative works on the RhB photodegradation performance of TiO2-based heterojunction photocatalysts.
Table 2. Summary of recent relative works on the RhB photodegradation performance of TiO2-based heterojunction photocatalysts.
PhotocatalystC0
(mg/L)
Dosage
(mg)
Light SourceDegradation RateTime
(min)
Kinetic Rate (min−1)Ref.
Ag@TiO210100150 W
Xe lamp
98.2%1200.0188[46]
TiO2 hollow boxes10050Visible light96.5%2400.0025[47]
Ag2O/TiO24.7940UV light87.7%800.0277[48]
Ag/ZnO/AgO/TiO21030350 W
Xe lamp
99.3%1000.0230[49]
Pt/A/R-TiO2--UV light92.4%900.0280[50]
Bi2WO6/TiO2/Pt20100UV light60.0%400.0210[51]
g-C3N4/TiO2505Visible light87.0%3000.0115[52]
A/R-TiO21025UV lightAbout 100%50-[53]
Au/A/R-TiO2--UV light97%600.0470[54]
TiN@(A,R)TiO21030Visible light97.0%900.0227This work

3.3. Photocatalytic Mechanism

Figure 7a shows the UV–Vis DRS spectra of all prepared catalysts. TiN shows full spectrum absorption characteristics similar to those of metals. Compared to P25 and sample S4 with little TiN content showing no obvious visible light absorption, the other samples show obvious light absorption in the entire visible light region (390–780 nm). Moreover, with increasing hydrolysis degree of TiN, the light absorption intensity gradually decreases, confirming that component TiN plays a decisive role in the light absorption ability of the prepared photocatalyst. The result is consistent with the report that the presence of TiN contributed to improving the material’s entire solar light absorption capability [31].
The band gap (Eg) is further obtained through the conversion of Formula (1) [55]:
αhν = A (hν − Eg) n⁄2,
where A is a constant, n = 1 for indirect semiconductors [56], and α and h are the absorption coefficient and photon energy, respectively.
With the decrease in the TiN content, the Eg of the sample gradually increases from 0.70 eV (S1) to 3.06 eV (P25). Therefore, the presence of TiN can effectively reduce the band gap of the sample, thus significantly improving the capability of light absorbance and utilization.
Figure 7c shows the PL spectral analysis of the samples, in which a higher fluorescence intensity represents a higher carrier recombination rate [57]. TiN and P25 showed the lowest and strongest fluorescence intensity, respectively. The fluorescence intensity of the others gradually increased with increasing TiO2 phase content. In particular, sample S2 also maintains low fluorescence intensity, indicating its outstanding charge separation ability. From the instantaneous photocurrent results in Figure 7d, the highest transient photocurrent signal can be observed for sample S2. The photocurrent signal decreases remarkably when the TiO2 phase content is further increased. The above results demonstrate the photocurrent enhancement effect of TiN on TiO2.
Figure 7e shows the electrochemical impedance spectroscopy (EIS) analysis of the samples. In addition, the resultant Nyquist plots (insert in Figure 7e) were fitted with an equivalent circuit using Zman software. As is shown, the equivalent circuit consists of internal resistance (Rs), charge transfer resistance (Rct1, Rct2), Warburg impedance (W), and double-layer capacitance (CPE1, CPE2) [58]. Compared to P25, the charge transfer resistance of the other samples (Figure 7e) is reduced to a certain extent, suggesting that the presence of TiN could improve the conductivity of the samples [59]. In particular, the charge transfer resistance of sample S2 is the lowest, demonstrating the greatest charge transfer rate. The above results demonstrate that the best charge carries separation and transfer can be obtained in sample S2. The reason can be explained by its proper phase proportion, and the formation of the N–O bond at the interface of (A,R)-TiO2 and TiN, which can effectively reduce the interface contact resistance of the heterostructure.
The calculated flat band potentials (Efb) are shown in Figure S5. The Mott–Schottky curves and calculation process of Efb can be found in Figures S5 and S6. Considering that the ECB of the n-type semiconductor is about 0.1 eV higher than its Efb, the ECB of the samples can be further deduced [60]. Combined with the band gap values (Eg), their energy band structures can be obtained (Figure 7f) through the following formula (2) [61]:
EVB = ECB + Eg.
It can be seen in Figure 7f that the presence of TiN can significantly decrease the Eg of the samples by improving the valence band (VB) potential. For sample S2, the Eg was reduced from 3.06 eV to 1.42 eV with the VB position changing from +2.04 to +0.82 eV, and CB position was slightly changed compared to P25. With a narrow band gap, sample S2 is more conducive to generating e and h+ charge carriers, while it shows no ability to produce ·OH active species due to its high valence band position (VB), which is in good agreement with the result of the EPR test.
Considering the staggered band structures of the A-TiO2/R-TiO2, Type-Ⅱ or Z-type charge transfer modes may occur in the heterojunction, as shown in Figure 8. The values of the CB and VB for (A,R)-TiO2 are obtained from the literature [62]. If Type-II mode is formed, electrons transfer to the CB of A-TiO2, the reduction potential of which is weak and cannot further reduce surface-adsorbed oxygen to generate ·O2 for the following photodegradation process, and this situation is inconsistent with our experimental results. Therefore, the Z-type charge transfer pathway is preferred for the heterojunction constructed in our work. Specifically, according to the literature, for partially reduced samples, the VB and CB positions of R-TiO2 are higher than that of A-TiO2 and the work function of R-TiO2 (φ ≈ 4.3 eV) is smaller than that of A-TiO2 (φ ≈ 4.7 eV) [63,64]. When they are in contact, free electrons spontaneously flow from R-TiO2 to A-TiO2 to obtain their Fermi energy levels to reach equilibrium. At this time, there are a large number of negatively charged electrons near the A-TiO2 interface. In contrast, positive charges are gathered at the R-TiO2 interface, generating a built-in electric field. Due to the shift in the Fermi energy level, R-TiO2 will generate an upward band bending, while A-TiO2 will generate a downward band bending [65]. The Z-type electron transfer path is generated due to the formed electric field and the energy band bending. To further prove the formed heterojunction is a Z-type photocatalyst, the Ag nanoparticles were photo-deposited on the catalyst to track where the electrons flow to. Figure 9 shows the EDS, TEM, and HRTEM images of the photodeposition of Ag nanoparticles on sample S2. It shows the uniform distribution of Ag, N, O, and Ti elements, and the Ag nanoparticles were isolated on R-TiO2 and apart from A-TiO2. The results suggest that the electrons were left on R-TiO2, confirming the Z-type charge transfer pathway in the formed heterojunction.
In this work, the plasmonic component TiN can broaden the absorbed light range and generate hot electrons due to the LSPR effect. Since nanostructured TiO2 was obtained in situ from TiN and charge transfer channel N–O bonds were formed between TiN and TiO2, the resulting intimate contacted interface benefits the electron transfer between them. Furthermore, the work function of TiN is ~3.7 eV (φm), and the electron affinity of TiO2 is ~4.2 eV (φs). Considering the barrier energy (the lowest energy required for an electron in the metal to be injected into the semiconductor) can be calculated as φ = φm − φs [66], a negative value (−0.5 eV) can be obtained, suggesting the quick injecting of the hot electrons into TiO2. Therefore, the improved photocatalytic performance of TiN@A-TiO2/R-TiO2 heterojunction can be concluded and shown in Figure 10. First, the plasmonic properties of TiN greatly broaden the light absorption range, generating and injecting hot electrons into TiO2. Furthermore, the N–O bond contacted TiO2/TiN heterointerface can significantly reduce the contact resistance of the interface and improve the charge transfer efficiency. Moreover, the optimized three-phase ratio and the formed Z-type A-TiO2/R-TiO2 heterojunction with an intimate interface contribute to the charge carrier separation and retain its high redox capacity. Thus, more active species will participate in the following photodegradation activities.

4. Conclusions

In this work, a plasmonic effect-enhanced TiN@A-TiO2/R-TiO2 direct Z-type heterojunction was fabricated through the simple hydrothermal reaction process. By regulating the amount of H2O2 oxidant, the proportion of TiN, anatase TiO2, and rutile TiO2 contents can be successfully adjusted and the interface charge transfer channel (N–O bond) has been constructed. Due to the Z-type charge transfer path between A-TiO2 and R-TiO2, the N–O bond connected charge transfer channel at the TiN/TiO2 interface, and the synergistic plasma resonance effect of TiN, the optimized photocatalyst shows a distinct increment in visible light absorption, photocurrent generation, and photocatalytic performance, demonstrating an effective approach to promote the photoactive properties of semiconductor photocatalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13131984/s1, Figure S1: SEM images of (a) TiN, (b) S1, (c) S2, (d) S3, (e) S4. Figure S2: Particle size distribution of all samples (a) TiN, (b) S1, (c) S2, (d) S3, (e) S4. Figure S3: N2 adsorption/desorption isotherm plots and pore size distribution of all samples. Figure S4: Free radical capture experiment of sample S2. Figure S5: Mott-Schottky curves of samples (a) P25, (b) TiN, (c) S1, (d) S2, (e) S3, (f) S4. Figure S6: Schematic diagram of flat band potential of the prepared samples.

Author Contributions

Conceptualization, C.L. (Changqing Liu) and C.L. (Chengxin Li); Data curation, W.W. and L.C.; Formal analysis, C.X.; Funding acquisition, Y.W. and C.L. (Changqing Liu); Investigation, W.W.; Methodology, L.C.; Project administration, Y.W. and C.L. (Chengxin Li); Resources, Y.W. and C.L. (Chengxin Li); Supervision, Y.W. and C.L. (Changqing Liu); Validation, C.L. (Changqing Liu); Writing—original draft, W.W.; Writing—review & editing, C.L. (Changqing Liu). All authors have read and agreed to the published version of the manuscript.

Funding

This work has been supported by the National Natural Science Foundation of China (Grant No. 51702194 and 52173214), the Natural Science Foundation of Shaanxi Province (Grant No. 2023-JC-YB-384), and the Youth Innovation Team of Shaanxi Universities (2022-70).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Masudy-Panah, S.; Siavash Moakhar, R.; Chua, C.S.; Kushwaha, A.; Dalapati, G.K. Stable and Efficient CuO Based Photocathode through Oxygen-Rich Composition and Au–Pd Nanostructure Incorporation for Solar-Hydrogen Production. ACS Appl. Mater. Interfaces 2017, 9, 27596–27606. [Google Scholar] [CrossRef] [PubMed]
  2. Chong, M.N.; Jin, B.; Chow, C.W.K.; Saint, C. Recent Developments in Photocatalytic Water Treatment Technology: A Review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef] [PubMed]
  3. Takata, T.; Domen, K. Particulate Photocatalysts for Water Splitting: Recent Advances and Future Prospects. ACS Energy Lett. 2019, 4, 542–549. [Google Scholar] [CrossRef]
  4. Tong, H.; Ouyang, S.; Bi, Y.; Umezawa, N.; Oshikiri, M.; Ye, J. Nano-Photocatalytic Materials: Possibilities and Challenges. Adv. Mater. 2012, 24, 229–251. [Google Scholar] [CrossRef]
  5. Hunge, Y.M.; Yadav, A.A.; Kang, S.-W.; Mohite, B.M. Role of Nanotechnology in Photocatalysis Application. Recent Pat. Nanotechnol. 2023, 17, 5–7. [Google Scholar] [CrossRef]
  6. Yadav, A.A.; Hunge, Y.M.; Kang, S.-W.; Fujishima, A.; Terashima, C. Enhanced Photocatalytic Degradation Activity Using the V2O5/RGO Composite. Nanomaterials 2023, 13, 338. [Google Scholar] [CrossRef]
  7. Carey, J.H.; Lawrence, J.; Tosine, H.M. Photodechlorination of PCB’s in the Presence of Titanium Dioxide in Aqueous Suspensions. Bull. Environ. Contam. Toxicol. 1976, 16, 697–701. [Google Scholar] [CrossRef]
  8. Poudel, M.B.; Kim, A.A. Silver Nanoparticles Decorated TiO2 Nanoflakes for Antibacterial Properties. Inorg. Chem. Commun. 2023, 152, 110675. [Google Scholar] [CrossRef]
  9. Hou, J.; Yang, C.; Wang, Z.; Jiao, S.; Zhu, H. Bi2O3 Quantum Dots Decorated Anatase TiO2 Nanocrystals with Exposed {001} Facets on Graphene Sheets for Enhanced Visible-Light Photocatalytic Performance. Appl. Catal. B Environ. 2013, 129, 333–341. [Google Scholar] [CrossRef]
  10. Kment, S.; Riboni, F.; Pausova, S.; Wang, L.; Wang, L.; Han, H.; Hubicka, Z.; Krysa, J.; Schmuki, P.; Zboril, R. Photoanodes Based on TiO2 and α-Fe2O3 for Solar Water Splitting—Superior Role of 1D Nanoarchitectures and of Combined Heterostructures. Chem. Soc. Rev. 2017, 46, 3716–3769. [Google Scholar] [CrossRef]
  11. Reddy, N.L.; Kumar, S.; Krishnan, V.; Sathish, M.; Shankar, M.V. Multifunctional Cu/Ag Quantum Dots on TiO2 Nanotubes as Highly Efficient Photocatalysts for Enhanced Solar Hydrogen Evolution. J. Catal. 2017, 350, 226–239. [Google Scholar] [CrossRef]
  12. Zhang, L.; Ding, N.; Lou, L.; Iwasaki, K.; Wu, H.; Luo, Y.; Li, D.; Nakata, K.; Fujishima, A.; Meng, Q. Localized Surface Plasmon Resonance Enhanced Photocatalytic Hydrogen Evolution via Pt@Au NRs/C3N4 Nanotubes under Visible-Light Irradiation. Adv. Funct. Mater. 2019, 29, 1806774. [Google Scholar] [CrossRef]
  13. Boerigter, C.; Campana, R.; Morabito, M.; Linic, S. Evidence and Implications of Direct Charge Excitation as the Dominant Mechanism in Plasmon-Mediated Photocatalysis. Nat. Commun. 2016, 7, 10545. [Google Scholar] [CrossRef] [Green Version]
  14. Guler, U.; Kildishev, A.V.; Boltasseva, A.; Shalaev, V.M. Plasmonics on the Slope of Enlightenment: The Role of Transition Metal Nitrides. Faraday Discuss. 2015, 178, 71–86. [Google Scholar] [CrossRef]
  15. Lima, L.P.B.; Diniz, J.A.; Doi, I.; Godoy Fo, J. Titanium Nitride as Electrode for MOS Technology and Schottky Diode: Alternative Extraction Method of Titanium Nitride Work Function. Microelectron. Eng. 2012, 92, 86–90. [Google Scholar] [CrossRef]
  16. Al-Hamdi, A.M.; Rinner, U.; Sillanpää, M. Tin Dioxide as a Photocatalyst for Water Treatment: A Review. Process Saf. Environ. Prot. 2017, 107, 190–205. [Google Scholar] [CrossRef]
  17. Mascaretti, L.; Barman, T.; Bricchi, B.R.; Münz, F.; Li Bassi, A.; Kment, Š.; Naldoni, A. Controlling the Plasmonic Properties of Titanium Nitride Thin Films by Radiofrequency Substrate Biasing in Magnetron Sputtering. Appl. Surf. Sci. 2021, 554, 149543. [Google Scholar] [CrossRef]
  18. Fakhouri, H.; Arefi-Khonsari, F.; Jaiswal, A.K.; Pulpytel, J. Enhanced Visible Light Photoactivity and Charge Separation in TiO2/TiN Bilayer Thin Films. Appl. Catal. Gen. 2015, 492, 83–92. [Google Scholar] [CrossRef] [Green Version]
  19. Clatworthy, E.B.; Yick, S.; Murdock, A.T.; Allison, M.C.; Bendavid, A.; Masters, A.F.; Maschmeyer, T. Enhanced Photocatalytic Hydrogen Evolution with TiO2 –TiN Nanoparticle Composites. J. Phys. Chem. C 2019, 123, 3740–3749. [Google Scholar] [CrossRef]
  20. Cherevan, A.S.; Gebhardt, P.; Shearer, C.J.; Matsukawa, M.; Domen, K.; Eder, D. Interface Engineering in Nanocarbon–Ta2O5 Hybrid Photocatalysts. Energy Env. Sci 2014, 7, 791–796. [Google Scholar] [CrossRef]
  21. Bai, S.; Ge, J.; Wang, L.; Gong, M.; Deng, M.; Kong, Q.; Song, L.; Jiang, J.; Zhang, Q.; Luo, Y.; et al. A Unique Semiconductor-Metal-Graphene Stack Design to Harness Charge Flow for Photocatalysis. Adv. Mater. 2014, 26, 5689–5695. [Google Scholar] [CrossRef]
  22. Maarisetty, D.; Baral, S.S. Defect Engineering in Photocatalysis: Formation, Chemistry, Optoelectronics, and Interface Studies. J. Mater. Chem. A 2020, 8, 18560–18604. [Google Scholar] [CrossRef]
  23. Zhu, H.; Yang, Y.; Kang, Y.; Niu, P.; Kang, X.; Yang, Z.; Ye, H.; Liu, G. Strong Interface Contact between NaYF4:Yb,Er and CdS Promoting Photocatalytic Hydrogen Evolution of NaYF4:Yb,Er/CdS Composites. J. Mater. Sci. Technol. 2022, 102, 1–7. [Google Scholar] [CrossRef]
  24. Li, J.; Gao, L.; Sun, J.; Zhang, Q.; Guo, J.; Yan, D. Synthesis of Nanocrystalline Titanium Nitride Powders by Direct Nitridation of Titanium Oxide. J. Am. Ceram. Soc. 2001, 84, 3045–3047. [Google Scholar] [CrossRef]
  25. Zhang, Y.; Huang, Y.; Zhu, S.; Liu, Y.; Zhang, X.; Wang, J.; Braun, A. Covalent S-O Bonding Enables Enhanced Photoelectrochemical Performance of Cu2S/Fe2O3 Heterojunction for Water Splitting. Small 2021, 17, 2100320. [Google Scholar] [CrossRef] [PubMed]
  26. Ning, Q.; Zhang, L.; Liu, C.; Li, X.; Xu, C.; Hou, X. Boosting Photogenerated Carriers for Organic Pollutant Degradation via In-Situ Constructing Atom-to-Atom TiO2/ZrTiO4 Heterointerface. Ceram. Int. 2021, 47, 33298–33308. [Google Scholar] [CrossRef]
  27. Guan, Y.; Hu, S.; Li, P.; Zhao, Y.; Wang, F.; Kang, X. In-Situ Synthesis of Highly Efficient Direct Z-Scheme Cu3P/g-C3N4 Heterojunction Photocatalyst for N2 Photofixation. Nano 2019, 14, 1950083. [Google Scholar] [CrossRef]
  28. Liu, X.; Zhang, Q.; Ma, D. Advances in 2D/2D Z-Scheme Heterojunctions for Photocatalytic Applications. Sol. RRL 2021, 5, 2000397. [Google Scholar] [CrossRef]
  29. Xu, Q.; Zhang, L.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Jaroniec, M. Direct Z-Scheme Photocatalysts: Principles, Synthesis, and Applications. Mater. Today 2018, 21, 1042–1063. [Google Scholar] [CrossRef]
  30. Liu, C.; Li, X.; Xu, C.; Wu, Y.; Hu, X.; Hou, X. Boron-Doped Rutile TiO2/Anatase TiO2/ZrTiO4 Ternary Heterojunction Photocatalyst with Optimized Phase Interface and Band Structure. Ceram. Int. 2020, 46, 20943–20953. [Google Scholar] [CrossRef]
  31. Kaur, M.; Shinde, S.L.; Ishii, S.; Jevasuwan, W.; Fukata, N.; Yu, M.-W.; Li, Y.; Ye, J.; Nagao, T. Marimo-Bead-Supported Core–Shell Nanocomposites of Titanium Nitride and Chromium-Doped Titanium Dioxide as a Highly Efficient Water-Floatable Green Photocatalyst. ACS Appl. Mater. Interfaces 2020, 12, 31327–31339. [Google Scholar] [CrossRef] [PubMed]
  32. Zhou, X.; Peng, F.; Wang, H.; Yu, H.; Yang, J. Preparation of Nitrogen Doped TiO2 Photocatalyst by Oxidation of Titanium Nitride with H2O2. Mater. Res. Bull. 2011, 46, 840–844. [Google Scholar] [CrossRef]
  33. Huang, D.G.; Liao, S.J.; Zhou, W.B.; Quan, S.Q.; Liu, L.; He, Z.J.; Wan, J.B. Synthesis of Samarium- and Nitrogen-Co-Doped TiO2 by Modified Hydrothermal Method and Its Photocatalytic Performance for the Degradation of 4-Chlorophenol. J. Phys. Chem. Solids 2009, 70, 853–859. [Google Scholar] [CrossRef]
  34. 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. 2021, 269, 116170. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, C.; Xu, C.; Wang, W.; Chen, L.; Li, X.; Wu, Y. Oxygen Vacancy Mediated Band-Gap Engineering via B-Doping for Enhancing Z-Scheme A-TiO2/R-TiO2 Heterojunction Photocatalytic Performance. Nanomaterials 2023, 13, 794. [Google Scholar] [CrossRef] [PubMed]
  36. Al-Dhaifallah, M.; Abdelkareem, M.A.; Rezk, H.; Alhumade, H.; Nassef, A.M.; Olabi, A.G. Co-decorated Reduced Graphene/Titanium Nitride Composite as an Active Oxygen Reduction Reaction Catalyst with Superior Stability. Int. J. Energy Res. 2021, 45, 1587–1598. [Google Scholar] [CrossRef]
  37. Wang, W.-K.; Chen, J.-J.; Gao, M.; Huang, Y.-X.; Zhang, X.; Yu, H.-Q. Photocatalytic Degradation of Atrazine by Boron-Doped TiO2 with a Tunable Rutile/Anatase Ratio. Appl. Catal. B Environ. 2016, 195, 69–76. [Google Scholar] [CrossRef]
  38. Zhang, X.; Zhang, Y.; Yu, Z.; Wei, X.; Wu, W.D.; Wang, X.; Wu, Z. Facile Synthesis of Mesoporous Anatase/Rutile/Hematite Triple Heterojunctions for Superior Heterogeneous Photo-Fenton Catalysis. Appl. Catal. B Environ. 2020, 263, 118335. [Google Scholar] [CrossRef]
  39. Hashemizadeh, I.; Golovko, V.B.; Choi, J.; Tsang, D.C.W.; Yip, A.C.K. Photocatalytic Reduction of CO2 to Hydrocarbons Using Bio-Templated Porous TiO2 Architectures under UV and Visible Light. Chem. Eng. J. 2018, 347, 64–73. [Google Scholar] [CrossRef]
  40. Hao, H.; Shi, J.-L.; Xu, H.; Li, X.; Lang, X. N-Hydroxyphthalimide-TiO2 Complex Visible Light Photocatalysis. Appl. Catal. B Environ. 2019, 246, 149–155. [Google Scholar] [CrossRef]
  41. Cheng, X.; Yu, X.; Xing, Z.; Yang, L. Synthesis and Characterization of N-Doped TiO2 and Its Enhanced Visible-Light Photocatalytic Activity. Arab. J. Chem. 2016, 9, S1706–S1711. [Google Scholar] [CrossRef]
  42. Sarkar, A.; Khan, G.G. The Formation and Detection Techniques of Oxygen Vacancies in Titanium Oxide-Based Nanostructures. Nanoscale 2019, 11, 3414–3444. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, Y.; Chen, J.; Hua, L.; Li, S.; Zhang, X.; Sheng, W.; Cao, S. High Photocatalytic Activity of Hierarchical SiO2@C-Doped TiO2 Hollow Spheres in UV and Visible Light towards Degradation of Rhodamine B. J. Hazard. Mater. 2017, 340, 309–318. [Google Scholar] [CrossRef]
  44. Poudel, M.B.; Awasthi, G.P.; Kim, H.J. Novel Insight into the Adsorption of Cr(VI) and Pb(II) Ions by MOF Derived Co-Al Layered Double Hydroxide @hematite Nanorods on 3D Porous Carbon Nanofiber Network. Chem. Eng. J. 2021, 417, 129312. [Google Scholar] [CrossRef]
  45. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, F.; Cheng, Z.; Cui, L.; Duan, T.; Anan, A.; Zhang, C.; Kang, L. Controllable Synthesis of Ag@TiO2 Heterostructures with Enhanced Photocatalytic Activities under UV and Visible Excitation. RSC Adv. 2016, 6, 1844–1850. [Google Scholar] [CrossRef]
  47. Zhao, X.; Du, Y.; Zhang, C.; Tian, L.; Li, X.; Deng, K.; Chen, L.; Duan, Y.; Lv, K. Enhanced Visible Photocatalytic Activity of TiO2 Hollow Boxes Modified by Methionine for RhB Degradation and NO Oxidation. Chin. J. Catal. 2018, 39, 736–746. [Google Scholar] [CrossRef]
  48. Liu, G.; Wang, G.; Hu, Z.; Su, Y.; Zhao, L. Ag2O Nanoparticles Decorated TiO2 Nanofibers as a p-n Heterojunction for Enhanced Photocatalytic Decomposition of RhB under Visible Light Irradiation. Appl. Surf. Sci. 2019, 465, 902–910. [Google Scholar] [CrossRef]
  49. Bian, H.; Zhang, Z.; Xu, X.; Gao, Y.; Wang, T. Photocatalytic Activity of Ag/ZnO/AgO/TiO2 Composite. Phys. E Low-Dimens. Syst. Nanostructures 2020, 124, 114236. [Google Scholar] [CrossRef]
  50. Wang, W.-K.; Chen, J.-J.; Zhang, X.; Huang, Y.-X.; Li, W.-W.; Yu, H.-Q. Self-Induced Synthesis of Phase-Junction TiO2 with a Tailored Rutile to Anatase Ratio below Phase Transition Temperature. Sci. Rep. 2016, 6, 20491. [Google Scholar] [CrossRef] [Green Version]
  51. Lu, Y.; Zhao, K.; Zhao, Y.; Zhu, S.; Yuan, X.; Huo, M.; Zhang, Y.; Qiu, Y. Bi2WO6/TiO2/Pt Nanojunction System: A UV–Vis Light Responsive Photocatalyst with High Photocatalytic Performance. Colloids Surf. Physicochem. Eng. Asp. 2015, 481, 252–260. [Google Scholar] [CrossRef]
  52. Li, Y.; Lv, K.; Ho, W.; Dong, F.; Wu, X.; Xia, Y. Hybridization of Rutile TiO2 (R-TiO2) with g-C3N4 Quantum Dots (CN QDs): An Efficient Visible-Light-Driven Z-Scheme Hybridized Photocatalyst. Appl. Catal. B Environ. 2017, 202, 611–619. [Google Scholar] [CrossRef]
  53. Zhang, X.; Lin, Y.; He, D.; Zhang, J.; Fan, Z.; Xie, T. Interface Junction at Anatase/Rutile in Mixed-Phase TiO2: Formation and Photo-Generated Charge Carriers Properties. Chem. Phys. Lett. 2011, 504, 71–75. [Google Scholar] [CrossRef]
  54. Yu, Y.; Wen, W.; Qian, X.-Y.; Liu, J.-B.; Wu, J.-M. UV and Visible Light Photocatalytic Activity of Au/TiO2 Nanoforests with Anatase/Rutile Phase Junctions and Controlled Au Locations. Sci. Rep. 2017, 7, 41253. [Google Scholar] [CrossRef] [Green Version]
  55. Xu, X.; Ding, X.; Yang, X.; Wang, P.; Li, S.; Lu, Z.; Chen, H. Oxygen Vacancy Boosted Photocatalytic Decomposition of Ciprofloxacin over Bi2MoO6: Oxygen Vacancy Engineering, Biotoxicity Evaluation and Mechanism Study. J. Hazard. Mater. 2019, 364, 691–699. [Google Scholar] [CrossRef]
  56. Baran, T.; Wojtyła, S.; Minguzzi, A.; Rondinini, S.; Vertova, A. Achieving Efficient H2O2 Production by a Visible-Light Absorbing, Highly Stable Photosensitized TiO2. Appl. Catal. B Environ. 2019, 244, 303–312. [Google Scholar] [CrossRef]
  57. Lakshmanareddy, N.; Navakoteswara Rao, V.; Cheralathan, K.K.; Subramaniam, E.P.; Shankar, M.V. Pt/TiO2 Nanotube Photocatalyst—Effect of Synthesis Methods on Valance State of Pt and Its Influence on Hydrogen Production and Dye Degradation. J. Colloid Interface Sci. 2019, 538, 83–98. [Google Scholar] [CrossRef]
  58. Poudel, M.B.; Kim, A.A.; Lohani, P.C.; Yoo, D.J.; Kim, H.J. Assembling Zinc Cobalt Hydroxide/Ternary Sulfides Heterostructure and Iron Oxide Nanorods on Three-Dimensional Hollow Porous Carbon Nanofiber as High Energy Density Hybrid Supercapacitor. J. Energy Storage 2023, 60, 106713. [Google Scholar] [CrossRef]
  59. Liu, Y.; Shen, S.; Zhang, J.; Zhong, W.; Huang, X. Cu2−xSe/CdS Composite Photocatalyst with Enhanced Visible Light Photocatalysis Activity. Appl. Surf. Sci. 2019, 478, 762–769. [Google Scholar] [CrossRef]
  60. Yin, W.; Bai, L.; Zhu, Y.; Zhong, S.; Zhao, L.; Li, Z.; Bai, S. Embedding Metal in the Interface of a p-n Heterojunction with a Stack Design for Superior Z-Scheme Photocatalytic Hydrogen Evolution. ACS Appl. Mater. Interfaces 2016, 8, 23133–23142. [Google Scholar] [CrossRef]
  61. Zhang, S.; Liu, Y.; Ma, R.; Jia, D.; Wen, T.; Ai, Y.; Zhao, G.; Fang, F.; Hu, B.; Wang, X. Molybdenum (VI)-oxo Clusters Incorporation Activates g-C3N4 with Simultaneously Regulating Charge Transfer and Reaction Centers for Boosting Photocatalytic Performance. Adv. Funct. Mater. 2022, 32, 2204175. [Google Scholar] [CrossRef]
  62. Guan, Y.; Liu, W.; Zuo, S.; Yuan, K.; Wu, F.; Ji, J.; Yao, C. Double Z-Scheme TiO2 (R)/C-TiO2 (A) Heterojunction Greatly Enhanced Efficiency of Photocatalytic Desulfurization under Sunlight. J. Mater. Sci. Mater. Electron. 2020, 31, 22297–22311. [Google Scholar] [CrossRef]
  63. Lin, Y.; Jiang, Z.; Zhu, C.; Hu, X.; Zhang, X.; Zhu, H.; Fan, J.; Lin, S.H. C/B Codoping Effect on Band Gap Narrowing and Optical Performance of TiO2 Photocatalyst: A Spin-Polarized DFT Study. J. Mater. Chem. A 2013, 1, 4516. [Google Scholar] [CrossRef]
  64. Kashiwaya, S.; Morasch, J.; Streibel, V.; Toupance, T.; Jaegermann, W.; Klein, A. The Work Function of TiO2. Surfaces 2018, 1, 73–89. [Google Scholar] [CrossRef] [Green Version]
  65. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-Scheme Heterojunction Photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  66. Li, C.; Yang, W.; Liu, L.; Sun, W.; Li, Q. In Situ Growth of TiO2 on TiN Nanoparticles for Non-Noble-Metal Plasmonic Photocatalysis. RSC Adv. 2016, 6, 72659–72669. [Google Scholar] [CrossRef]
Figure 1. XRD results of the prepared samples obtained with various H2O2 content.
Figure 1. XRD results of the prepared samples obtained with various H2O2 content.
Nanomaterials 13 01984 g001
Figure 2. FT-IR spectra of TiN and sample S2.
Figure 2. FT-IR spectra of TiN and sample S2.
Nanomaterials 13 01984 g002
Figure 3. (a) TEM; (b) HRTEM; (c) HADDF images and EDS mappings of the elements N (d); O (e); Ti (f) for S2.
Figure 3. (a) TEM; (b) HRTEM; (c) HADDF images and EDS mappings of the elements N (d); O (e); Ti (f) for S2.
Nanomaterials 13 01984 g003
Figure 4. XPS spectra of S2: (a) full spectrum; (b) Ti 2p; and (c) O1s.
Figure 4. XPS spectra of S2: (a) full spectrum; (b) Ti 2p; and (c) O1s.
Nanomaterials 13 01984 g004
Figure 5. (a) Photodegradation performance; (b) kinetics of all prepared catalysts; (c) cycling experiments of sample S2; and (d) XRD patterns of the as-prepared and cycled sample S2.
Figure 5. (a) Photodegradation performance; (b) kinetics of all prepared catalysts; (c) cycling experiments of sample S2; and (d) XRD patterns of the as-prepared and cycled sample S2.
Nanomaterials 13 01984 g005
Figure 6. EPR results of (a) DMPO-•O2; (b) DMPO-•OH with sample S2.
Figure 6. EPR results of (a) DMPO-•O2; (b) DMPO-•OH with sample S2.
Nanomaterials 13 01984 g006
Figure 7. (a) UV–Vis DRS; (b) Plot of (αhv)1/2 versus hν; (c) PL spectra; (d) TP curves; (e) EIS plots; (f) band structures of the samples S1 (blue), S2 (red), S3 (purple), S4 (orange), P25 (yellowish-brown) and TiN (green).
Figure 7. (a) UV–Vis DRS; (b) Plot of (αhv)1/2 versus hν; (c) PL spectra; (d) TP curves; (e) EIS plots; (f) band structures of the samples S1 (blue), S2 (red), S3 (purple), S4 (orange), P25 (yellowish-brown) and TiN (green).
Nanomaterials 13 01984 g007
Figure 8. Schematic illustration of the possible charge carrier transfer mode in Type-Ⅱ and direct Z-type photocatalysts.
Figure 8. Schematic illustration of the possible charge carrier transfer mode in Type-Ⅱ and direct Z-type photocatalysts.
Nanomaterials 13 01984 g008
Figure 9. (a) TEM; (b) HRTEM; (c) HADDF images and EDS mappings of the elements Ag (d); N (e); O (f); Ti (g) for S2 with photo-deposited Ag nanoparticles.
Figure 9. (a) TEM; (b) HRTEM; (c) HADDF images and EDS mappings of the elements Ag (d); N (e); O (f); Ti (g) for S2 with photo-deposited Ag nanoparticles.
Nanomaterials 13 01984 g009
Figure 10. (a,b) Schematic illustration of the formation and a possible photoinduced catalytic mechanism of the TiN@(A,R)TiO2 heterojunction.
Figure 10. (a,b) Schematic illustration of the formation and a possible photoinduced catalytic mechanism of the TiN@(A,R)TiO2 heterojunction.
Nanomaterials 13 01984 g010
Table 1. Compositions and ratios of the A/R-TiO2 and TiN in the prepared samples.
Table 1. Compositions and ratios of the A/R-TiO2 and TiN in the prepared samples.
SamplesContent of H2O2 (wt%)TiO2/wt%TiN/wt%
A-TiO2R-TiO2A-TiO2:R-TiO2
TiN0---100%
S10.580.9%19.1%4.2521.1%
S21.093.8%6.2%15.137.2%
S32.597.3%2.7%36.401.7%
S45.098.3%1.7%57.820.2%
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

Wang, W.; Wu, Y.; Chen, L.; Xu, C.; Liu, C.; Li, C. Fabrication of Z-Type TiN@(A,R)TiO2 Plasmonic Photocatalyst with Enhanced Photocatalytic Activity. Nanomaterials 2023, 13, 1984. https://doi.org/10.3390/nano13131984

AMA Style

Wang W, Wu Y, Chen L, Xu C, Liu C, Li C. Fabrication of Z-Type TiN@(A,R)TiO2 Plasmonic Photocatalyst with Enhanced Photocatalytic Activity. Nanomaterials. 2023; 13(13):1984. https://doi.org/10.3390/nano13131984

Chicago/Turabian Style

Wang, Wanting, Yuanting Wu, Long Chen, Chenggang Xu, Changqing Liu, and Chengxin Li. 2023. "Fabrication of Z-Type TiN@(A,R)TiO2 Plasmonic Photocatalyst with Enhanced Photocatalytic Activity" Nanomaterials 13, no. 13: 1984. https://doi.org/10.3390/nano13131984

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