Next Article in Journal
Application of Zeolites and Zeolitic Imidazolate Frameworks in Dentistry—A Narrative Review
Next Article in Special Issue
Geometrical Stabilities and Electronic Structures of Rh5 Nanoclusters on Rutile TiO2 (110) for Green Hydrogen Production
Previous Article in Journal
The Impact of Ambient Temperature on Electrothermal Characteristics in Stacked Nanosheet Transistors with Multiple Lateral Stacks
Previous Article in Special Issue
Free Convection in a Square Ternary Hybrid Nanoliquid Chamber with Linearly Heating Adjacent Walls
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Catalytic Performance of Ag NP/0.95AgNbO3-0.05LiTaO3 Heterojunction from the Combination of Surface Plasma Resonance Effect and Piezoelectric Effect Using Facile Mechanical Milling

1
Chemical Engineering College, Inner Mongolia University of Technology, Hohhot 010051, China
2
Engineering Research Center of Large Energy Storage Technology, Ministry of Education, Inner Mongolia University of Technology, Hohhot 010051, China
3
Key Laboratory of Inorganic Function Material and Device, Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 200050, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(22), 2972; https://doi.org/10.3390/nano13222972
Submission received: 20 September 2023 / Revised: 4 November 2023 / Accepted: 15 November 2023 / Published: 18 November 2023

Abstract

:
An internal built electric field can suppress the recombination of electron–hole pairs and distinctly enhance the catalytic activity of a photocatalyst. Novel t-Ag/0.95AgNbO3-0.05LiTaO3 heterojunction was prepared by reducing silver nanoparticles (Ag NPs) on the surface of the piezoelectric powder 0.95AgNbO3-0.05LiTaO3 (0.05-ANLT) using a simple mechanical milling method. The effects of milling time and excitation source used for the degradation of organic dye by heterojunction catalysts were investigated. The results demonstrate that the optimized 1.5-Ag/0.05-ANLT heterojunction removes 97% RhB within 40 min, which is 7.8 times higher than that of single piezoelectric catalysis and 25.4 times higher than that of single photocatalysis. The significant enhancement of photocatalytic activity can be attributed to the synergistic coupling of the surface plasmon resonance (SPR) effect and the piezoelectric effect.

1. Introduction

The continuous utilization of fossil fuels results in serious environmental pollution and energy shortage, which blocks the sustainable development of contemporary society. Among these problems, water pollution has been further exacerbated by the extensive release of chemical dyes from the textile and printing sectors [1,2,3]. Photocatalysis excited by sunlight can decompose organic pollutants in water and provide a low-cost and environmental friendly solution [4,5,6]. However, the limited light response of conventional single-component photocatalysts and the high electron–hole pair recombination rate greatly hinder their quantum efficiency, thereby imposing significant constraints on their practical applications [7]. Many methods have been proposed to overcome these disadvantages [8,9,10,11]. Due to its ability to broaden the light absorption spectrum and mitigate electron–hole recombination, heterojunction has drawn more and more attention in recent years [12].
Unfortunately, the heterojunction can only drive the photoinduced charges near the junction region to take part in the photocatalysis reaction, which means that the electron and hole in the bulk of the semiconductor have been left to recombine. The piezopotential induced by external stress in piezoelectric has been proven to separate excited charges in deep regions [13]. For instance, zinc oxide (ZnO) nanowires reduced the recombination of electron–hole pairs under external pressure and enhanced the degradation of methylene blue (MB) [14].
It is hypothesized that the combination of heterojunction and piezopotential could further enhance the activity of the catalyst. Nevertheless, the piezo-photocatalysis performance of common piezoelectric is hindered by the low visible light response from their large band gap (>3 eV) [15]. Currently, heterostructure construction [16], defect engineering [17], and chemical modification [18] have been used to enhance the catalytic efficacy of piezoelectric materials. It is well known that the design of morphotropic phase boundaries (MPBs) greatly improved the electrical properties of piezoelectric materials [19]. The coexistence of multiple phases provides more directions for polarization, which facilitates polarization rotation. The piezoelectric potential is directly determined by the piezoelectric constant according to the following equation: V = d33 (SYL)/ε0ε, where d33 is the piezoelectric constant, L is the original thickness, Y is Young’s modulus of the piezoelectric, ε0 is the permittivity of free space, and ε is the relative dielectric constant. Exceptional piezo-photocatalytic performance has been observed near the morphotropic phase boundary in BiPrFeMnO3 nanofibers [20], Sm-doped PMN-PT [21], and KNN [22] powder.
AgNbO3, with a band gap of approximately 2.8 eV, shows notable visible light absorption [23]. The distinctive d10 electronic structure and plasmonic resonance effect of AgNbO3 make it a versatile catalyst for hydrogen production [24,25] and the degradation of organic pollutants [26,27]. On the other hand, strong ferroelectric and piezoelectric properties can be achieved via the formation of a solid solution between AgNbO3 and other ferroelectrics (Ag1−xKxNbO3 [28], (Ag1−xLix)NbO3 [29], and (1 − x)AgNbO3-xLiTaO3 [30]). In addition, the manipulation of its ferro/piezoelectric properties has been established as a significant approach to enhancing the photocatalytic efficacy of AgNbO3 [31,32].
The plasmon resonance effect (SPR) from nano noble metal (Au, Ag, Pt) particles on surfaces can produce a strong localized electromagnetic field and enhance the efficiency of electron and hole separation in semiconductors [33,34]. Larger K+ substitution for Ag+ transforms AgNbO3 from an antiferroelectric to a ferroelectric state, building a substantial internal electric field. Additionally, a minute quantity of metallic silver can be generated at elevated levels of K+ doping. Combining the SPR and piezoelectric potential, the piezo-photocatalytic degradation has also been sharply enhanced [35]. At present, the methods of loading noble metal on semiconductors mainly include piezoelectric electrochemical deposition [36], photochemical reduction [37], and impregnation [38]. Mechanochemical synthesis through ball milling can avoid the use of hazardous organic solvents and external heating, shorten reaction times, and simplify the synthesis process [39]. Silver nanoparticles (Ag NPs) were deposited onto TiO2 via mechanical ball milling. Compared with pure TiO2, the degradation rate of methyl orange (MO) dye under UV irradiation by Ag/TiO2 heterojunction has been sharply increased (2.1 times) [40].
Recently, superior catalytic activity (1 − x)AgNbO3-xLiTaO3 solid solution near the antiferroelectric–ferroelectric (AFE-FE) phase boundary has been reported by our group [41]. It was found that the color of the fresh solid solution (light yellow) has changed to dark yellow during dry milling, which indicates that some reaction has occurred. In this study, t-Ag/0.95AgNbO3-0.05LiTaO3 (hereafter referred to as t-Ag/0.05-ANLT) piezo-photocatalyst was synthesized through a facile mechanical grinding. The piezo-photocatalytic degradation of the organic dye of this heterojunction has been investigated. Within 40 min, the removal rate of RhB by 1.5-Ag/0.05-ANLT heterojunction under light and ultrasonic vibration was 97%, which was 7.8 times that of piezoelectric catalysis alone and 25.4 times that of photocatalysis alone.

2. Materials and Methods

2.1. Materials

Analytical-grade niobium pentoxide (Nb2O5, 99%), lithium carbonate (Li2CO3, 98%), silver oxide (Ag2O, 99.8%), and tantalum pentoxide (Ta2O5, 99.9%), along with ethanol and rhodamine B (RhB), were procured from Sinopharm Chemical Reagent Corp. (Shanghai, China). The chemicals were employed in their original form without additional purification. Furthermore, all aqueous solutions were prepared using deionized water.

2.2. Preparation of the Samples

In this study, 0.05-ANLT powder was synthesized using a traditional solid-state reaction technique. In this process, a thorough mixing of Ag2O, Nb2O5, Li2CO3, and Ta2O5 was achieved in anhydrous ethanol for 30 min using a mortar. The resulting powders were then dried and calcined at a temperature of 900 °C for 6 h while maintaining a constant oxygen flow rate of 5 mL/min.
After that, a grinding process was used to form Ag NPs on the surface of 0.05-ANLT. Specifically, 0.2 g of 0.05-ANLT solid solution powder was subjected to grinding with zirconia balls in a mini-mill (MSK-SFM-12M, Hefei Kejing, Hefei, China), with the sample-to-ball mass ratio of 1:2. The 0.05-ANLT powders ground for 0.5, 1, 1.5, and 2 h were denoted as 0.5-Ag/0.05-ANLT, 1-Ag/0.05-ANLT, 1.5-Ag/0.05-ANLT, and 2-Ag/0.05-ANLT, respectively. During the grinding process, it is imperative to shield the sample from light.

2.3. Characterization

Powder X-ray diffraction (PXRD) was utilized to examine the composition of the milled powder. The measurement was conducted using Cu Ka radiation (Rigaku Corporation, Tokyo, Japan) with a voltage of 45 kV and a current of 200 mA. Ultraviolet-visible spectroscopy (UV-vis) measurement was performed using a UV-vis spectrophotometer (Hitachi UV-3600, Tokyo, Japan) equipped with an integrating sphere attachment. To further investigate the morphology of the samples, transmission electron microscope (TEM) and high-resolution TEM (HRTEM) (Talos F200X, FEI, Thermo Fisher Scientific, Waltham, MA, USA) were employed. The photoluminescence (PL) spectra were acquired via fluorescence spectroscopy (RF-5301PC, Shimadzu, Tokyo, Japan) at a stimulation wavelength of 365 nm.

2.4. Piezo-Photocatalytic Characterization

The piezo-photocatalytic activity of the t-Ag/0.05-ANLT catalyst was assessed through the degradation of RhB. This evaluation involved the utilization of 300 W Xenon lamp (CEL-HXF300, CEAULIGHT, Beijing, China), which was equipped with a UV cutoff filter to produce visible light above 420 nm. An ultrasonic cleaner (40 kHz and 300 W, KQ-300DE, Kunshan, China) was employed as an ultrasonic source. In this study, 100 mg of t-Ag/0.05-ANLT powder was dispersed into 100 mL RhB dye solution of 5 mg/L. To attain adsorption–desorption equilibrium, the mixture was agitated in the absence of light for one hour. Throughout the degradation process, approximately 3 mL of the solution was extracted in 10 min intervals and subsequently filtered through a 0.45-μm Millipore filter. The concentration of RhB was determined using a UV-Vis spectrometer (UVT6, Beijing Purkinje General Instrument Co., Ltd., Beijing, China) by measuring its characteristic absorbance at 554 nm. The suspension subjected to ultrasonic vibration was maintained at room temperature to prevent any heat-induced catalytic effects.

2.5. Detection of Active Species in Catalysis

To unveil the active species engaged in the degradation of RhB dye, radical capture experiments were conducted. Three distinct scavengers for holes (h+), superoxide ions (•O2), and hydroxyl radicals (•OH), namely triethanolamine (TEOA), benzoquinone (BQ), and isopropyl alcohol (IPA) were employed correspondingly.

2.6. Electrochemical Performance Test

To measure the photocurrent, electrochemical impedance (EIS) and Mott–Schottky curve, we employed a conventional three-compartment cell consisting of a working electrode, a Pt wire counter electrode, and a reference electrode saturated with calomel. The electrolyte solution was 0.1 M Na2SO4. Xenon lamp light source (Zolix HPS-300XA, Beijing, China) was used to irradiate the electrochemical cell. The current and voltage signals were obtained using the I-t program of the electrochemical workstation (Chi660E, Chenhua, Shanghai, China). EIS measurements were conducted with an AC amplitude of 5 mV in a frequency range of 10−2 to 105 Hz. The Mott–Schottky curves of the working electrode were recorded at frequencies of 1000 Hz using the same electrochemical workstation (Chi660E, Chenhua, Shanghai, China).

3. Results

Figure 1 displays the X-ray diffraction (XRD) patterns of t-Ag NP/0.05-ANLT catalysts. Similar diffraction patterns have been observed for all powders after grinding for different times. These diffraction peaks match well with the standard card of 0.05-ANLT solid solution (JCPDS No. 53-0346) [30]. This finding confirms the high crystallinity of t-Ag/0.05-ANLT powders and indicates that the crystal structure of 0.05-ANLT solid solution remains unchanged basically after grinding. As reported in the literature, the powder of 0.05-ANLT is composed of orthorhombic (O) and rhombohedral (R) phases [35]. This composition has high piezoelectricity. However, the color of the powder is darkened, which indicates that some Ag might be reduced out [33]. The absence of diffraction peaks corresponding to Ag in the XRD pattern can be attributed to its extremely low concentration, which falls below the detection limit of X-ray. Further investigation is necessary to validate the presence of Ag on the surface of the 0.05-ANLT particle.
TEM and HRTEM results of 1.5-Ag/0.05-ANLT composite are shown in Figure 2. It is reported that a smooth and clean surface can be seen in fresh 0.05-ANLT. The size of the particles varies between 400 and 800 nm [41]. As seen from Figure 2a, the particle size decreases to 150–500 nm after mechanical grinding for 1.5 h. It should be stressed that some nanoparticles are uniformly dispersed on the surface of the large one. The lattice fringe widths of 0.234 nm and 0.277 nm (Figure 2b) correspond to the (111) plane of Ag (JCPDS No. 04-0783) and the (114) plane of AgNbO3 (JCPDS No. 52-0405), respectively. The size of Ag NPs is about 7 nm, and they are tightly anchored to the surface of 0.05-ANLT solid solution particle and form the heterojunction [42].
Generally, the noble metal nanoparticle on the semiconductors will change the light absorbance of the semiconductor through the surface plasma effect [43,44]. Figure 3a compares the DRS absorption spectra of 0.05-ANLT powder before and after grinding. Fresh 0.05-ANLT absorbs UV light up to 450 nm. However, the light absorption from 450 to 700 nm has been enhanced due to Ag nanoparticles on the surface. The increased intensity of the peak around 650 nm indicates that the Ag content rises [33]. In addition, with the increase in Ag content, SPR absorption is significantly enhanced. Generally speaking, the conduction band bottom of AgNbO3 is composed of O 2p and Nb 4d orbitals, but the Nb 4d orbital contributes more. Furthermore, the valence band of the AgNbO3 photocatalyst consists of O 2p and Ag 4d orbitals, which are motivated and generate electrons and holes under visible-light illumination. Hence, the excited state electrons mainly derive from the Nb 4d orbital of the conduction band [27]. The SPR effect from Ag NPs not only improves the visible light absorption of 0.05-ANLT but also separates photogenerated carriers, which might enhance visible light photocatalytic activity [33,45].
Furthermore, the band gaps (Eg) of 0.05-ANLT and t-Ag/0.05-ANLT are calculated by using the following equation [46]:
( A h υ ) 2 / n h υ E g
where A, , and Eg are the absorbance, the irradiation energy, and the band gap, respectively. In the context of direct and indirect semiconductors, the value of n is 1 for a direct semiconductor, while 4 is for an indirect semiconductor. As AgNbO3 belongs to a direct-gap semiconductor [27], the corresponding n value is 1 in present case. The Eg values of the samples with 0.05-ANLT are approximately 2.87 eV (±0.03), as depicted in Figure 3b. The band gap energies of 0.05-ANLT samples after grinding for different times of 0.5, 1, 1.5, and 2 h are 2.91 eV (±0.02), 2.94 eV (±0.03), 2.89 eV (±0.04), and 2.93 eV (±0.02). These findings suggest that the band gap width of 0.05-ANLT specimens exhibits negligible variation before and after grinding due to the extremely low content of Ag. Clearly, the alteration in band gap has minimal impact on the catalytic performance of the specimen.
The piezo-photocatalytic degradation performance of t-Ag/0.05-ANLT for RhB under visible light and ultrasonic vibration was investigated in Figure 4. The results showed that the degradation rate was negligible without a catalyst, even when ultrasonic vibration and simulated visible light irradiation were applied. These findings suggest that the t-Ag/0.05-ANLT catalyst is crucial for the degradation of RhB. The fresh 0.05-ANLT decomposes 60% RhB within 40 min. However, compared with pure 0.05-ANLT, all t-Ag/0.05-ANLT catalysts exhibited higher piezo-photocatalytic degradation rate. The piezo-photocatalytic degradation rates of 0.05-ANLT were 60%, 64%, 89%, 97%, and 68% after mechanical milling for 0.5, 1, 1.5, and 2 h, respectively. Among them, the 1.5-Ag/0.05-ANLT catalyst exhibits the best piezo-photocatalytic performance. Obviously, the t-Ag/0.05-ANLT heterojunction effectively enhances piezoelectric photocatalytic performance.
The photocatalytic degradation of RhB follows first-order kinetics [47]. The kinetic equation can be represented as follows:
ln ( C 0 / C ) = k t
where k denotes the apparent pseudo-first-order rate constant (min−1), C represents the concentration of the organic dye at time t (mol L−1), and C0 signifies the initial concentration of the organic dye (mol L−1). The degradation of the dye leads to the decoloration of the RhB solution. Figure 4b shows the degradation kinetic data, and the rate constants (k values) can be determined according to Equation (2). Figure 4c shows that the piezo-photocatalytic reaction rate constants of 0.05-ANLT after grinding for 0, 0.5, 1.0, 1.5, and 2.0 h are 0.02325 min−1, 0.02571 min−1, 0.54470 min−1, 0.08434 min−1 and 0.02772 min−1, respectively. The proper amount of Ag NPs on the surface enhances piezo-photocatalytic performance. However, further increase in Ag NPs may induce a shielding phenomenon, impeding the penetration of light radiation into 0.05-ANLT, leading to marked reduction in catalytic efficacy [48]. Figure 4d shows the absorbance spectra of RhB degraded by the 1.5-Ag/0.05-ANLT heterojunction. The gradual decrease in the absorption peak at 554 nm in Figure 4d indicates that RhB has been degraded.
Degradation experiments under single light irradiation, single vibration irradiation, and the co-excitation of light and ultrasonic vibration were conducted to unveil the synergetic effect of photocatalysis and piezocatalysis. Figure 5a shows the degradation efficiency of RhB dye by 1.5-Ag/0.05-ANLT sample. After visible light irradiation for 40 min, only 13% RhB has been decomposed. Within the same time, the degradation rate increased to 35% under ultrasonic vibration. However, most RhB (97%) could be degraded under coexcitation of light and vibration. Figure 5b compares their kinetic rate constants (k). The degradation constants (k) of photocatalysis and piezocatalysis for RhB were 0.00332 min−1 and 0.01082 min−1, respectively. However, the bicatalysis rate constant sharply increased to 0.08434 min−1. It is suggested that a strong synergetic effect is responsible for this distinct improvement, similar with Au/AgNbO3 [32] and Aux/BaTiO3 [49].
To examine the university of the degradation capability of the 1.5-Ag/0.05-ANLT catalyst, MB and MO solutions were also degraded by 1.5-Ag/0.05-ANLT. As illustrated in Figure 6a, the degradation efficiencies of RhB, MB, and MO reached 97%, 96%, and 89%, respectively, within 40 min. Compared with MO, 1.5-Ag/0.05-ANLT has slightly higher degradation efficiency for RhB and MB dyes because their negatively charged surfaces readily adsorb the cationic dyes [50]. The degradation rate constants shown in Figure 6b for RhB and MB are 0.08434 min−1 and 0.08311 min−1, which was slightly higher than that for MO (0.07521 min−1). The result indicates that 1.5-Ag/0.05-ANLT can quickly decompose different dyes and work as a promising piezo-photocatalyst.
The PL emission occurs when electron–hole pairs recombine in semiconductor. The higher PL intensity indicates the stronger carrier recombination and lower photocatalytic activity [51]. In this study, 0.05-ANLT and 1.5-Ag/0.05-ANLT exhibit similar emission bands from 400 to 650 nm in Figure 7a. The higher intensity of the PL peak for 0.05-ANLT indicates the stronger recombination of the electron and hole. By contrast, the separation of charge carriers has been improved in 1.5-Ag/0.05-ANLT. The charge separation in photocatalysts could also be characterized by its current under light irradiation [52]. In Figure 7b, the photocurrent of the 1.5-Ag/0.05-ANLT heterojunction is higher than that of the fresh 0.05-ANLT, confirming its enhanced charge carrier separation. To further elucidate the rationale behind the enhanced piezo-photocatalytic activity, Nyquist curves have been employed to evaluate the charge migration resistance. Generally, a smaller semicircle diameter signifies lower charge transfer resistance. Figure 7c suggests that the 1.5 h-Ag/0.05-ANLT composite exhibits a smaller semicircle diameter. The Mott–Schottky in Figure 7d demonstrates that all samples belong to n-type semiconductors [53]. The flat-band potential of the 1.5 h-Ag/0.05-ANLT composite at the x-axis intercept is −0.951 V, more negative than that of the 0.05-ANLT composite (−0.905 V). A higher flat band potential corresponds to the stronger reducible ability of 1.5 h-Ag/0.05-ANLT composite [54].
In general, the ECB for n-type semiconductors is 0.1–0.3 eV lower than the flat-band potential value [55,56,57,58]. As Figure 7d shows, ECB values for 0.05-ANLT and 1.5 h-Ag/0.05-ANLT are −0.805 (vs. NHE) and −0.851 V (vs. NHE), respectively. The Eg of 0.05-ANLT is approximately 2.87 eV (±0.03) in Figure 3b. The band gap of 0.05-ANLT samples after grinding for 1.5 h is 2.89 eV (±0.04). Based on the formula ECB = EVBEg [59], the EVB of 0.05-ANLT and 1.5 h-Ag/0.05-ANLT are 2.065 V and 2.039 V, respectively. These values are comparable with the literature [26,32].
The stability of the catalyst is very important in practical applications. The results in Figure 8 demonstrate that the t-Ag/0.05-ANLT piezo-photocatalyst still can degrade 88% RhB after four recycles. This result implies that the heterojunction shows acceptable stability. On the other hand, Yu et al. found that the XRD pattern of Ag/AgNbO3 by the combustion method before and after photoreduction does not change and concluded that this composite catalyst is stable in solution under light irradiation [60]. A similar process might happen in the present case, which indicates the notable stability of our catalyst.
To identify the active species in the piezo-photocatalytic reaction and reveal its reaction mechanism, a series of degradation experiments of RhB have been conducted with the addition of different scavengers. Figure 9 shows that the degradation of RhB is 97% in the absence of any scavengers. However, a great reduction in the degradation rate has been observed (35%, 28% and 14%) after the addition of IPA, BQ and TEOA. Since triethanolamine (TEOA) captures h+, it indicates that h+ functions as the most important active specie in the present case. However, •OH and •O2 also play important roles during piezo-photocatalysis of RhB.
Table 1 presents some typical catalytic degradation performances of AgNbO3-based composite catalysts for RhB. The reaction rate constant of t-Ag/0.05-ANLT piezo-photocatalyst surpasses other AgNbO3-based photocatalytic materials.
According to the active species experiment, the potential piezo-photocatalytic mechanism of t-Ag/0.05-ANLT can be proposed in Figure 10. AgNbO3 is an n-type semiconductor (conduction band bottom −0.805 V; valence band top 2.045 V) [26,62]. Ag has a higher work function (4.62 eV) than AgNbO3 (4.485 eV) [31,63]. When Ag NPs are coupled with 0.05-ANLT, the Schottky barrier forms at the metal-semiconductor interface. Upon illumination, visible light causes a collective oscillation of electrons within the Ag NPs on the surface of the t-Ag/0.05-ANLT heterojunction, which enhances visible light absorption (Figure 3) [60]. The hot electrons with high energy surpass the Schottky barrier, migrate to the conduction band of t-Ag/0.05-ANLT and leave holes in Ag Nps, suppressing the recombination of electrons and holes [32]. In addition, the piezoelectric potential under ultrasonic vibration tilts the conduction and valence bands of the 0.05-ANLT catalyst and drives electrons and holes to the opposite direction [64]. This modification, in turn, decreases the barrier height for hot electrons to transition to the semiconductor (Figure 10b). The photogenerated e and h+ in the t-Ag/0.05-ANLT solid solution are effectively separated [23]. After that, electrons on the conduction band of t-Ag/0.05-ANLT reduces O2 to •O2, while h+ on the top of the valance band forms highly active •OH [65]. h+, •O2 and •OH effectively degrade the adsorbed organic dyes. Additionally, Ag NPs transfer the hot electrons to 0.05-ANLT, and thus the induced charge carriers are separated efficiently. Simultaneously, Ag NPs also act as a “fast channel”, facilitating the electron migration to the Ag/dye solution interface. All these greatly boost the generation of reactive radicals and enhance the photocatalytic performance [48]. Hence, the notable augmentation in the catalytic performance of 1.5-Ag/0.05-ANLT heterostructure should be ascribed to the contribution of the SPR effect of Ag NP and the piezoelectric potential of 0.05-ANLT.

4. Conclusions

Novel t-Ag NP/0.05-ANLT composite was successfully constructed using a facile mechanical milling method. Milling time has an important impact on the catalytic performance of a heterostructure. The optimized composition decomposes 97% RhB within 40 min. The SPR effect from Ag NP enhances visible light absorption of heterojunction. The decomposition activity of the optimal composition under the coexcitaion of ultrasonic and visible light has been increased by 6.8 and 24.4 times than bare piezocatalysis and photocatalysis. The injection of hot electrons from Ag NP and piezopotential from piezoelectric 0.05-ANLT greatly promotes the separation of photoinduced electron and hole, which are responsible for the distinct enhancement of catalytic activity.

Author Contributions

Conceptualization, Z.C. and G.L.; methodology, T.R. and T.H.; validation T.H.; formal analysis T.R.; resources P.X.; data curation X.T.; writing—original draft preparation, T.R.; writing—review and editing, Z.C.; visualization, Z.C.; supervision, Z.C.; project administration, G.L.; funding acquisition, Z.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Science Foundation of Inner Mongolia Autonomous Regions (No. 2023MS05015), Program for universities directedly under Inner Mongolia Autonomous Regions (No. JY20230095), Program for Grassland Elite (No. CYYC10032), Program for Innovative Research Team in Universities of Inner Mongolia Autonomous Region (No. NMGIRT2214).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mishra, S.; Cheng, L.; Maiti, A. The utilization of agro-biomass/byproducts for effective bio-removal of dyes from dyeing wastewater: A comprehensive review. J. Environ. Chem. Eng. 2021, 9, 104901. [Google Scholar] [CrossRef]
  2. Uddin, F. Environmental hazard in textile dyeing wastewater from local textile industry. Cellulose 2021, 28, 10715–10739. [Google Scholar] [CrossRef]
  3. Liang, J.; Ning, X.; Sun, J.; Song, J.; Hong, Y.; Cai, H. An integrated permanganate and ozone process for the treatment of textile dyeing wastewater: Efficiency and mechanism. J. Clean. Prod. 2018, 204, 12–19. [Google Scholar] [CrossRef]
  4. Yagub, M.T.; Sen, T.K.; Afroze, S.; Ang, H.M. Dye and its removal from aqueous solution by adsorption: A review. Adv. Colloid. Interface Sci. 2014, 209, 172–184. [Google Scholar] [CrossRef] [PubMed]
  5. Wu, N. Plasmonic metal-semiconductor photocatalysts and photoelectrochemical cells: A review. Nanoscale 2018, 10, 2679–2696. [Google Scholar] [CrossRef]
  6. Li, L.; Salvador, P.A.; Rohrer, G.S. Photocatalysts with internal electric fields. Nanoscale 2014, 6, 24–42. [Google Scholar] [CrossRef]
  7. Zhang, L.; Mohamed, H.H.; Dillert, R.; Bahnemann, D. Kinetics and mechanisms of charge transfer processes in photocatalytic systems: A review. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 263–276. [Google Scholar] [CrossRef]
  8. Yu, C.; He, J.; Tan, M.; Hou, Y.; Zeng, H.; Liu, C.; Meng, H.; Su, Y.; Qiao, L.; Lookman, T.; et al. Selective Enhancement of Photo-Piezocatalytic Performance in BaTiO3 Via heterovalent Ion Doping. Adv. Funct. Mater. 2022, 32, 2209365. [Google Scholar] [CrossRef]
  9. Chen, C.; Ma, W.; Zhao, J. Semiconductor-mediated photodegradation of pollutants under visible-light irradiation. Chem. Soc. Rev. 2010, 39, 4206–4219. [Google Scholar] [CrossRef] [PubMed]
  10. Bansal, J.; Hafiz, A.K.; Sharma, S.N. Photoreduction of Dye with Noble Metal Gold Permeated with Metal Oxide Titania. J. Nanosci. Nanotechnol. 2020, 20, 3896–3901. [Google Scholar] [CrossRef]
  11. Liu, Q.; Zhai, D.; Xiao, Z.; Tang, C.; Sun, Q.; Bowen, C.R.; Luo, H.; Zhang, D. Piezo-photoelectronic coupling effect of BaTiO3@TiO2 nanowires for highly concentrated dye degradation. Nano Energy 2022, 92, 106702. [Google Scholar] [CrossRef]
  12. Deng, F.; Zhang, Q.; Yang, L.; Luo, X.; Wang, A.; Luo, S.; Dionysiou, D.D. Visible-light-responsive graphene-functionalized Bi-bridge Z-scheme black BiOCl/Bi2O3 heterojunction with oxygen vacancy and multiple charge transfer channels for efficient photocatalytic degradation of 2-nitrophenol and industrial wastewater treatment. Appl. Catal. B Environ. 2018, 238, 61–69. [Google Scholar] [CrossRef]
  13. Wang, P.; Zhong, S.; Lin, M.; Lin, C.; Lin, T.; Gao, M.; Zhao, C.; Li, X.; Wu, X. Signally enhanced piezo-photocatalysis of Bi0.5Na0.5TiO3/MWCNTs composite for degradation of rhodamine B. Chemosphere 2022, 308, 136596. [Google Scholar] [CrossRef]
  14. Xue, X.; Zang, W.; Deng, P.; Wang, Q.; Xing, L.; Zhang, Y.; Wang, Z.L. Piezo-potential enhanced photocatalytic degradation of organic dye using ZnO nanowires. Nano Energy 2015, 13, 414–422. [Google Scholar] [CrossRef]
  15. Zhao, W.; Zhang, Q.; Wang, H.; Rong, J.; Lei, E.; Dai, Y. Enhanced catalytic performance of Ag2O/BaTiO3 heterostructure microspheres by the piezo/pyro-phototronic synergistic effect. Nano Energy 2020, 73, 104783. [Google Scholar] [CrossRef]
  16. Hong, D.; Zang, W.; Guo, X.; Fu, Y.; He, H.; Sun, J.; Xing, L.; Liu, B.; Xue, X. High Piezo-photocatalytic Efficiency of CuS/ZnO Nanowires Using Both Solar and Mechanical Energy for Degrading Organic Dye. ACS Appl. Mater. Interfaces 2016, 8, 21302–21314. [Google Scholar] [CrossRef]
  17. Lin, L.; Feng, X.; Lan, D.; Chen, Y.; Zhong, Q.; Liu, C.; Cheng, Y.; Qi, R.; Ge, J.; Yu, C.; et al. Coupling Effect of Au Nanoparticles with the Oxygen Vacancies of TiO2–x for Enhanced Charge Transfer. J. Phys. Chem. C 2020, 124, 23823–23831. [Google Scholar] [CrossRef]
  18. Zhou, X.; Yan, F.; Wu, S.; Shen, B.; Zeng, H.; Zhai, J. Remarkable Piezophoto Coupling Catalysis Behavior of BiOX/BaTiO3 (X = Cl, Br, Cl(0.166) Br(0.834)) Piezoelectric Composites. Small 2020, 16, e2001573. [Google Scholar] [CrossRef]
  19. Dong, W.; Xiao, H.; Jia, Y.; Chen, L.; Geng, H.; Bakhtiar, S.U.H.; Fu, Q.; Guo, Y. Engineering the Defects and Microstructures in Ferroelectrics for Enhanced/Novel Properties: An Emerging Way to Cope with Energy Crisis and Environmental Pollution. Adv. Sci. 2022, 9, e2105368. [Google Scholar] [CrossRef] [PubMed]
  20. You, D.; Liu, L.; Yang, Z.; Xing, X.; Li, K.; Mai, W.; Guo, T.; Xiao, G.; Xu, C. Polarization-induced internal electric field to manipulate piezo-photocatalytic and ferro-photoelectrochemical performance in bismuth ferrite nanofibers. Nano Energy 2022, 93, 106852. [Google Scholar] [CrossRef]
  21. Yuan, B.; Wu, J.; Qin, N.; Lin, E.; Kang, Z.; Bao, D. Sm-doped Pb(Mg1/3Nb2/3)O3-xPbTiO3 piezocatalyst: Exploring the relationship between piezoelectric property and piezocatalytic activity. Appl. Mater. Today 2019, 17, 183–192. [Google Scholar] [CrossRef]
  22. Zhang, A.; Liu, Z.; Xie, B.; Lu, J.; Guo, K.; Ke, S.; Shu, L.; Fan, H. Vibration catalysis of eco-friendly Na0.5K0.5NbO3-based piezoelectric: An efficient phase boundary catalyst. Appl. Catal. B Environ. 2020, 279, 119353. [Google Scholar] [CrossRef]
  23. Li, L.; Ma, Y.; Chen, G.; Wang, J.; Wang, C. Oxygen-vacancy-enhanced piezo-photocatalytic performance of AgNbO3. Scr. Mater. 2022, 206, 114234. [Google Scholar] [CrossRef]
  24. Zhao, W.; Ai, Z.; Zhu, X.; Zhang, M.; Shi, Q.; Dai, J. Visible-light-driven photocatalytic H2 evolution from water splitting with band structure tunable solid solution (AgNbO3)1−x(SrTiO3)x. Int. J. Hydrogen Energy 2014, 39, 7705–7712. [Google Scholar] [CrossRef]
  25. Liu, J.; Zhang, S.; Jia, Y.; Tie, M.; Fang, D.; Zhang, Z.; Wang, J. Fabrication of novel immobilized and forced Z-scheme Ag|AgNbO3/Ag/Er3+:YAlO3@Nb2O5 nanocomposite film photocatalyst for enhanced degradation of auramine O with synchronous evolution of pure hydrogen. Sep. Purif. Technol. 2022, 288, 120658. [Google Scholar] [CrossRef]
  26. Wang, C.; Yan, J.; Wu, X.; Song, Y.; Cai, G.; Xu, H.; Zhu, J.; Li, H. Synthesis and characterization of AgBr/AgNbO3 composite with enhanced visible-light photocatalytic activity. Appl. Surf. Sci. 2013, 273, 159–166. [Google Scholar] [CrossRef]
  27. Yang, M.; Pu, Y.; Wang, W.; Li, J.; Guo, X.; Shi, R.; Shi, Y. Highly efficient Ag2O/AgNbO3 p-n heterojunction photocatalysts with enhanced visible-light responsive activity. J. Alloys Compd. 2019, 811, 151831. [Google Scholar] [CrossRef]
  28. Fu, D.; Itoh, M.; Koshihara, S.-Y. Dielectric, ferroelectric, and piezoelectric behaviors of AgNbO3–KNbO3 solid solution. J. Appl. Phys. 2009, 106, 104104. [Google Scholar] [CrossRef]
  29. Song, A.; Wang, J.; Song, J.; Zhang, J.; Li, Z.; Zhao, L. Antiferroelectricity and ferroelectricity in A-site doped silver niobate lead-free ceramics. J. Eur. Ceram. Soc. 2021, 41, 1236–1243. [Google Scholar] [CrossRef]
  30. Li, S.; Nie, H.; Wang, G.; Liu, N.; Zhou, M.; Cao, F.; Dong, X. Novel AgNbO3-based lead-free ceramics featuring excellent pyroelectric properties for infrared detecting and energy-harvesting applications via antiferroelectric/ferroelectric phase-boundary design. J. Mater. Chem. C 2019, 7, 4403–4414. [Google Scholar] [CrossRef]
  31. Wang, F.; Song, J.; Wang, T.; Du, C.; Su, Y. Photosynergy of Ag In Situ Anchored on AgNb1–xTaxO3 Solid Solutions as an Efficient and Durable Catalyst toward Nitrobenzene Reduction: Uncovering the Relevance of the Electronic Structure, Active Sites, and Catalytic Activity. J. Phys. Chem. C 2020, 125, 385–395. [Google Scholar] [CrossRef]
  32. Li, S.; Zhao, Z.; Liu, M.; Liu, X.; Huang, W.; Sun, S.; Jiang, Y.; Liu, Y.; Zhang, J.; Zhang, Z. Remarkably enhanced photocatalytic performance of Au/AgNbO3 heterostructures by coupling piezotronic with plasmonic effects. Nano Energy 2022, 95, 107031. [Google Scholar] [CrossRef]
  33. Lu, Y.; Shen, Q.; Yu, Q.; Zhang, F.; Li, G.; Zhang, W. Photoinduced In Situ Growth of Ag Nanoparticles on AgNbO3. J. Phys. Chem. C 2016, 120, 28712–28716. [Google Scholar] [CrossRef]
  34. Wang, P.; Huang, B.; Qin, X.; Zhang, X.; Dai, Y.; Wei, J.; Whangbo, M.H. Ag@AgCl: A highly efficient and stable photocatalyst active under visible light. Angew. Chem. Int. Ed. Engl. 2008, 47, 7931–7933. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, Y.; Luo, N.; Zeng, D.; Xu, C.; Ma, L.; Luo, G.; Qian, Y.; Feng, Q.; Chen, X.; Hu, C.; et al. Ferroelectricity and Schottky Heterojunction Engineering in AgNbO3: A Simultaneous Way of Boosting Piezo-photocatalytic Activity. ACS Appl. Mater. Interfaces 2022, 14, 22313–22323. [Google Scholar] [CrossRef] [PubMed]
  36. Lin, E.; Kang, Z.; Wu, J.; Huang, R.; Qin, N.; Bao, D. BaTiO3 nanocubes/cuboids with selectively deposited Ag nanoparticles: Efficient piezocatalytic degradation and mechanism. Appl. Catal. B Environ. 2021, 285, 119823. [Google Scholar] [CrossRef]
  37. Lin, E.; Wu, J.; Qin, N.; Yuan, B.; Bao, D. Silver modified barium titanate as a highly efficient piezocatalyst. Catal. Sci. Technol. 2018, 8, 4788–4796. [Google Scholar] [CrossRef]
  38. Chen, Z.; Zhou, H.; Kong, F.; Wang, M. Piezocatalytic oxidation of 5-hydroxymethylfurfural to 5-formyl-2-furancarboxylic acid over Pt decorated hydroxyapatite. Appl. Catal. B Environ. 2022, 309, 121281. [Google Scholar] [CrossRef]
  39. Yu, F.; Tian, F.; Zou, H.; Ye, Z.; Peng, C.; Huang, J.; Zheng, Y.; Zhang, Y.; Yang, Y.; Wei, X.; et al. ZnO/biochar nanocomposites via solvent free ball milling for enhanced adsorption and photocatalytic degradation of methylene blue. J. Hazard. Mater. 2021, 415, 125511. [Google Scholar] [CrossRef] [PubMed]
  40. Aysin, B.; Ozturk, A.; Park, J. Silver-loaded TiO2 powders prepared through mechanical ball milling. Ceram. Int. 2013, 39, 7119–7126. [Google Scholar] [CrossRef]
  41. He, T.; Cao, Z.; Li, G.; Jia, Y.; Peng, B. High efficiently harvesting visible light and vibration energy in (1−x)AgNbO3–xLiTaO3 solid solution around antiferroelectric–ferroelectric phase boundary for dye degradation. J. Adv. Ceram. 2022, 11, 1641–1653. [Google Scholar] [CrossRef]
  42. Zhang, Z.; Ji, Y.; Li, J.; Zhong, Z.; Su, F. Synergistic effect in bimetallic copper–silver (CuxAg) nanoparticles enhances silicon conversion in Rochow reaction. RSC Adv. 2015, 5, 54364–54371. [Google Scholar] [CrossRef]
  43. Lan, J.; Zhou, X.; Liu, G.; Yu, J.; Zhang, J.; Zhi, L.; Nie, G. Enhancing photocatalytic activity of one-dimensional KNbO3 nanowires by Au nanoparticles under ultraviolet and visible-light. Nanoscale 2011, 3, 5161–5167. [Google Scholar] [CrossRef]
  44. Wan, J.; Liu, E.; Fan, J.; Hu, X.; Sun, L.; Tang, C.; Yin, Y.; Li, H.; Hu, Y. In-situ synthesis of plasmonic Ag/Ag3PO4 tetrahedron with exposed {111} facets for high visible-light photocatalytic activity and stability. Ceram. Int. 2015, 41, 6933–6940. [Google Scholar] [CrossRef]
  45. Xiang, D.; Liu, Z.; Wu, M.; Liu, H.; Zhang, X.; Wang, Z.; Wang, Z.L.; Li, L. Enhanced Piezo-Photoelectric Catalysis with Oriented Carrier Migration in Asymmetric Au-ZnO Nanorod Array. Small 2020, 16, e1907603. [Google Scholar] [CrossRef]
  46. Yang, L.; Liu, J.; Chang, H.; Tang, S. Enhancing the visible-light-induced photocatalytic activity of AgNbO3 by loading Ag@AgCl nanoparticles. RSC Adv. 2015, 5, 59970–59975. [Google Scholar] [CrossRef]
  47. Konstantinou, I.K.; Albanis, T.A. TiO2-assisted photocatalytic degradation of azo dyes in aqueous solution: Kinetic and mechanistic investigations. Appl. Catal. B Environ. 2004, 49, 1–14. [Google Scholar] [CrossRef]
  48. Jiang, X.; Wang, H.; Wang, X.; Yuan, G. Synergetic effect of piezoelectricity and Ag deposition on photocatalytic performance of barium titanate perovskite. Sol. Energy 2021, 224, 455–461. [Google Scholar] [CrossRef]
  49. Xu, S.; Guo, L.; Sun, Q.; Wang, Z.L. Piezotronic Effect Enhanced Plasmonic Photocatalysis by AuNPs/BaTiO3 Heterostructures. Adv. Funct. Mater. 2019, 29, 1808737. [Google Scholar] [CrossRef]
  50. Zhou, X.; Sun, Q.; Zhai, D.; Xue, G.; Luo, H.; Zhang, D. Excellent catalytic performance of molten-salt-synthesized Bi0.5Na0.5TiO3 nanorods by the piezo-phototronic coupling effect. Nano Energy 2021, 84, 105936. [Google Scholar] [CrossRef]
  51. Zhang, S.; Zhang, B.-p.; Li, S.; Huang, Z.; Yang, C.; Wang, H. Enhanced photocatalytic activity in Ag-nanoparticle-dispersed BaTiO3 composite thin films: Role of charge transfer. J. Adv. Ceram. 2017, 6, 1–10. [Google Scholar] [CrossRef]
  52. Zhang, D.; Li, J.; Wang, Q.; Wu, Q. High {001} facets dominated BiOBr lamellas: Facile hydrolysis preparation and selective visible-light photocatalytic activity. J. Mater. Chem. A 2013, 1, 8622–8629. [Google Scholar] [CrossRef]
  53. Zheng, J.; Lei, Z. Incorporation of CoO nanoparticles in 3D marigold flower-like hierarchical architecture MnCo2O4 for highly boosting solar light photo-oxidation and reduction ability. Appl. Catal. B Environ. 2018, 237, 1–8. [Google Scholar] [CrossRef]
  54. Pan, H.; Heagy, M.D. Photons to Formate: A Review on Photocatalytic Reduction of CO2 to Formic Acid. Nanomaterials 2020, 10, 2422. [Google Scholar] [CrossRef] [PubMed]
  55. Zong, X.; Yan, H.; Wu, G.; Ma, G.; Wen, F.; Wang, L.; Li, C. Enhancement of photocatalytic H2 evolution on CdS by loading MoS2 as cocatalyst under visible light irradiation. J. Am. Chem. Soc. 2008, 130, 7176–7177. [Google Scholar] [CrossRef]
  56. Jang, J.S.; Kim, H.G.; Lee, J.S. Heterojunction semiconductors: A strategy to develop efficient photocatalytic materials for visible light water splitting. Catal. Today 2012, 185, 270–277. [Google Scholar] [CrossRef]
  57. Chen, H.; Leng, W.; Xu, Y. Enhanced visible-light photoactivity of CuWO4 through a surface-deposited CuO. J. Phys. Chem. C 2014, 118, 9982–9989. [Google Scholar] [CrossRef]
  58. Pinaud, B.A.; Chen, Z.; Abram, D.N.; Jaramillo, T.F. Thin films of sodium birnessite-type MnO2: Optical properties, electronic band structure, and solar photoelectrochemistry. J. Phys. Chem. C 2011, 115, 11830–11838. [Google Scholar] [CrossRef]
  59. Wen, X.-J.; Zhang, C.; Niu, C.-G.; Zhang, L.; Zeng, G.-M.; Zhang, X.-G. Highly enhanced visible light photocatalytic activity of CeO2 through fabricating a novel p–n junction BiOBr/CeO2. Catal. Commun. 2017, 90, 51–55. [Google Scholar] [CrossRef]
  60. Yu, Z.; Zhan, B.; Ge, B.; Zhu, Y.; Dai, Y.; Zhou, G.; Yu, F.; Wang, P.; Huang, B.; Zhan, J. Synthesis of high efficient and stable plasmonic photocatalyst Ag/AgNbO3 with specific exposed crystal-facets and intimate heterogeneous interface via combustion route. Appl. Surf. Sci. 2019, 488, 485–493. [Google Scholar] [CrossRef]
  61. Li, G.; Bai, Y.; Zhang, W.F.; Zhang, H. Enhanced visible light photocatalytic properties of AgNbO3/AgSbO3 composites. Mater. Chem. Phys. 2013, 139, 1009–1013. [Google Scholar] [CrossRef]
  62. Wu, W.; Liang, S.; Chen, Y.; Shen, L.; Yuan, R.; Wu, L. Mechanism and improvement of the visible light photocatalysis of organic pollutants over microcrystalline AgNbO3 prepared by a sol–gel method. Mater. Res. Bull. 2013, 48, 1618–1626. [Google Scholar] [CrossRef]
  63. Liang, Y.; Wang, C.; Kei, C.; Hsueh, Y.; Cho, W.; Perng, T. Photocatalysis of Ag-Loaded TiO2 Nanotube Arrays Formed by Atomic Layer Deposition. J. Phys. Chem. C 2011, 115, 9498–9502. [Google Scholar] [CrossRef]
  64. Li, S.; Zhao, Z.; Zhao, J.; Zhang, Z.; Li, X.; Zhang, J. Recent Advances of Ferro-, Piezo-, and Pyroelectric Nanomaterials for Catalytic Applications. ACS Appl. Nano Mater. 2020, 3, 1063–1079. [Google Scholar] [CrossRef]
  65. Chen, L.; Dai, X.; Li, X.; Wang, J.; Chen, H.; Hu, X.; Lin, H.; He, Y.; Wu, Y.; Fan, M. A novel Bi2S3/KTa0.75Nb0.25O3 nanocomposite with high efficiency for photocatalytic and piezocatalytic N2 fixation. J. Mater. Chem. A 2021, 9, 13344–13354. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of t-Ag NP/0.05-ANLT samples after grinding for different times.
Figure 1. XRD patterns of t-Ag NP/0.05-ANLT samples after grinding for different times.
Nanomaterials 13 02972 g001
Figure 2. TEM and HRTEM images of 1.5-Ag/0.05-ANLT (a,b).
Figure 2. TEM and HRTEM images of 1.5-Ag/0.05-ANLT (a,b).
Nanomaterials 13 02972 g002
Figure 3. (a) UV-vis absorption spectra and (b) the estimated band gaps of 0.05-ANLT samples at different grinding times.
Figure 3. (a) UV-vis absorption spectra and (b) the estimated band gaps of 0.05-ANLT samples at different grinding times.
Nanomaterials 13 02972 g003
Figure 4. (a) Degradation activity of t-Ag/0.05-ANLT piezo-photocatalysts under visible light and ultrasonic excitation. (b) Reaction kinetics curve of t-Ag/0.05-ANLT piezo-photocatalyst. (c) Degradation reaction rate constants of t-Ag/0.05-ANLT piezo-photocatalyst. (d) UV-vis spectral absorption of RhB for 1.5-Ag/0.05-ANLT under the irradiation of light and ultrasound.
Figure 4. (a) Degradation activity of t-Ag/0.05-ANLT piezo-photocatalysts under visible light and ultrasonic excitation. (b) Reaction kinetics curve of t-Ag/0.05-ANLT piezo-photocatalyst. (c) Degradation reaction rate constants of t-Ag/0.05-ANLT piezo-photocatalyst. (d) UV-vis spectral absorption of RhB for 1.5-Ag/0.05-ANLT under the irradiation of light and ultrasound.
Nanomaterials 13 02972 g004
Figure 5. (a) Comparison of catalytic activities for 1.5-Ag/0.05-ANLT under different excitations. (b) Degradation rate constants for 1.5-Ag/0.05-ANLT under different excitations.
Figure 5. (a) Comparison of catalytic activities for 1.5-Ag/0.05-ANLT under different excitations. (b) Degradation rate constants for 1.5-Ag/0.05-ANLT under different excitations.
Nanomaterials 13 02972 g005
Figure 6. (a) The piezo-photocatalytic activity of 1.5-Ag/0.05-ANLT for different dyes. (b) Corresponding degradation rate constants.
Figure 6. (a) The piezo-photocatalytic activity of 1.5-Ag/0.05-ANLT for different dyes. (b) Corresponding degradation rate constants.
Nanomaterials 13 02972 g006
Figure 7. (a) photoluminescence spectra, (b) photocurrent, (c) EIS curve, (d) Mott–Schottky plots for 0.05-ANLT and 1.5-Ag/0.05-ANLT.
Figure 7. (a) photoluminescence spectra, (b) photocurrent, (c) EIS curve, (d) Mott–Schottky plots for 0.05-ANLT and 1.5-Ag/0.05-ANLT.
Nanomaterials 13 02972 g007aNanomaterials 13 02972 g007b
Figure 8. Cycling experimental of piezo-photocatalytic degradation of RhB by 1.5-Ag/0.05-ANLT.
Figure 8. Cycling experimental of piezo-photocatalytic degradation of RhB by 1.5-Ag/0.05-ANLT.
Nanomaterials 13 02972 g008
Figure 9. Effect of active capture agents on the piezo-photocatalytic degradation of RhB by 1.5-Ag/0.05-ANLT.
Figure 9. Effect of active capture agents on the piezo-photocatalytic degradation of RhB by 1.5-Ag/0.05-ANLT.
Nanomaterials 13 02972 g009
Figure 10. Schematic illustration of the coupled plasmonic and piezo-photocatalytic process in the t-Ag/0.05-ANLT heterostructure. (a) t-Ag/0.05-ANLT under visible light irradiation; (b) t-Ag/0.05-ANLT under the excitation of visible light and ultrasonic.
Figure 10. Schematic illustration of the coupled plasmonic and piezo-photocatalytic process in the t-Ag/0.05-ANLT heterostructure. (a) t-Ag/0.05-ANLT under visible light irradiation; (b) t-Ag/0.05-ANLT under the excitation of visible light and ultrasonic.
Nanomaterials 13 02972 g010
Table 1. Typical photocatalytic performances of AgNbO3-based catalyst.
Table 1. Typical photocatalytic performances of AgNbO3-based catalyst.
Ferroelectric MaterialsPollutantsExcitation
Source
Concentration of Pollutantk × 10−3 (min−1)Ref.
Ag/AgNbO3RhBVisible light5.0 mg/L44.70[60]
Ag2O/AgNbO3RhBVisible light5.0 mg/L30.56[27]
AgNbO3/AgSbO3RhBVisible light2.5 mg/L46.66[61]
0.05-ANLTRhBVisible light + 300 W ultrsonic5.0 mg/L26.66[41]
t-Ag/0.05-ANLTRhBVisible light + 300 W ultrsonic5.0 mg/L84.34This work
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

Ren, T.; He, T.; Cao, Z.; Xing, P.; Teng, X.; Li, G. Enhanced Catalytic Performance of Ag NP/0.95AgNbO3-0.05LiTaO3 Heterojunction from the Combination of Surface Plasma Resonance Effect and Piezoelectric Effect Using Facile Mechanical Milling. Nanomaterials 2023, 13, 2972. https://doi.org/10.3390/nano13222972

AMA Style

Ren T, He T, Cao Z, Xing P, Teng X, Li G. Enhanced Catalytic Performance of Ag NP/0.95AgNbO3-0.05LiTaO3 Heterojunction from the Combination of Surface Plasma Resonance Effect and Piezoelectric Effect Using Facile Mechanical Milling. Nanomaterials. 2023; 13(22):2972. https://doi.org/10.3390/nano13222972

Chicago/Turabian Style

Ren, Tianxiang, Tufeng He, Zhenzhu Cao, Pengyue Xing, Xinglong Teng, and Guorong Li. 2023. "Enhanced Catalytic Performance of Ag NP/0.95AgNbO3-0.05LiTaO3 Heterojunction from the Combination of Surface Plasma Resonance Effect and Piezoelectric Effect Using Facile Mechanical Milling" Nanomaterials 13, no. 22: 2972. https://doi.org/10.3390/nano13222972

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