Next Article in Journal
Ultra-Thin Ion Exchange Membranes by Low Ionomer Blending for Energy Harvesting
Next Article in Special Issue
Facile Synthesis to Porous TiO2 Nanostructures at Low Temperature for Efficient Visible-Light Degradation of Tetracycline
Previous Article in Journal
Correction: Nguyen et al. Green Silver Nanoparticles Formed by Phyllanthus urinaria, Pouzolzia zeylanica, and Scoparia dulcis Leaf Extracts and the Antifungal Activity. Nanomaterials 2020, 10, 542
Previous Article in Special Issue
Recent Advances toward Enhanced Photocatalytic Proprieties of BiFeO3-Based Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

LDH/MXene Synergistic Carrier Separation Effects to Improve the Photoelectric Catalytic Activities of Bi2WO6 Nanosheet Arrays

State Key Laboratory of Heavy Oil Processing, China University of Petroleum, Beijing 102249, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2024, 14(5), 477; https://doi.org/10.3390/nano14050477
Submission received: 31 January 2024 / Revised: 23 February 2024 / Accepted: 5 March 2024 / Published: 6 March 2024
(This article belongs to the Special Issue Advanced Nanomaterials for Photocatalysis)

Abstract

:
Photoelectric catalysis is a green and efficient way to degrade pollutants, which has been paid more and more attention by researchers. Among them, Bi2WO3 has been proved to have excellent photocatalytic oxidation activity on its {001} facets. In this study, {001}-oriented facets with high exposure were successfully integrated into Bi2WO6 nanoplate arrays (Bi2WO6 NAs) to create a photoelectrode. This structure was grown in situ on an indium tin oxide (ITO) substrate. To promote photogenerated carrier separation efficiency and reduce agglomeration of Bi2WO6 photocatalysts, the electrochemical deposition of NiFe–layered double hydroxide (NiFe-LDH) and Ti3C2 (MXene) were introduced in this research to synergistically catalyze pollutant degradation. Morphology, spectral characterization, and electrochemical analysis jointly confirmed that the outstanding performance of hole capture behavior with LDH and electron conduction properties with MXene were the main reasons for the improvement in catalytic activity of the photoelectrode. Taking bisphenol A (BPA) as the model pollutant, the rate constant k of the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode reaches 0.00196 min−1 under photoelectrocatalytic (PEC) conditions, which is 4.5 times that of the pure Bi2WO6 NAs photoelectrode. This work provides a new way to improve the reaction kinetics of the PEC degradation of pollutants.

1. Introduction

Photoelectrocatalysis, as a green way to solve water pollution, has the advantages of electrocatalytic (EC) and photocatalytic (PC) oxidation [1,2]. During the photoelectrocatalytic (PEC) process, photogenerated electrons are driven towards the counter electrode by an external potential. This mechanism facilitates the efficient separation of photogenerated electrons and holes, thereby significantly enhancing photocatalytic activity [3,4,5,6].
As a significant semiconductor photocatalytic material, Bi2WO6 is the simplest compound in the Aurivillius family, which is alternately composed of [Bi2O2]2+ layers and perovskite-like (WO4)2− octahedral layers, and is widely used in PEC oxidation [3,7]. As a narrow band gap semiconductor (2.75 eV), Bi2WO6 has the advantages of a non-toxic, high-stability, and visible-light response [8,9,10]. Nevertheless, this material exhibits certain drawbacks, such as a high rate of electron and hole recombination, a low specific surface area, and a tendency towards easy agglomeration [4,11,12]. In order to address these issues, a variety of strategies have been employed to improve the photocatalytic performance of Bi2WO6. These methods include doping, the construction of heterojunctions, and surface hybridization [8,13,14,15]. In 2005, Zhu and co-workers synthesized Bi2WO6 nanostructures for the first time through simple hydrothermal treatment, which set a precedent for the preparation of Bi2WO6 nanostructures [16]. Cao and co-workers successfully prepared the 2D/2D heterojunction of ultra-thin Ti3C2/Bi2WO6 nanosheets, which have a short charge transfer distance and large interface contact area. Under simulated solar light irradiation, the combined yield of CH4 and CH3OH achieved by the Ti3C2/Bi2WO6 composite is 4.6 times higher than that of pure Bi2WO6 in the CO2 reduction reaction [17].
Layered double hydroxide (LDH) is a type of layered inorganic functional material, characterized by a positively charged layer and an anion that is exchangeable in the interlayer space [18,19,20]. LDH has attracted much attention due to its unique layered structure and exchangeable anion between layers [21,22]. In recent years, Zhu et al. reported three-dimensional BiVO4/NiFe-LDH heterostructures. The combination of BiVO4 and NiFe-LDH can accelerate photogenerated charge separation and photoelectrocatalytic activity [23]. Zhang and colleagues synthesized a ZnCr–CO3LDH/ruptured tubular g-C3N4 photocatalyst and observed that its capability to separate electron–hole pairs was enhanced by approximately fivefold and twofold, respectively, compared to bulk g-C3N4 and ruptured tubular g-C3N4 [24]. MXenes, a novel class of two-dimensional (2D) layered materials, are composed of transition metal carbides or carbonitrides. The general formula for MXenes is Mn+1Xn (where n = 1, 2, or 3), with ‘M’ representing a transition metal (such as Sc, Ti, Zr, Hf, V, Nb, Ta, or Mo), and ‘X’ denoting carbon (C) and/or nitrogen (N). Notably, MXenes (Ti3C2), synthesized by selectively etching out the ‘A’ layers from Ti3AC2 (where ‘A’ can be Al, Zn, Si, or Ga), have become a focal point in research on the synthesis of MXene-based photocatalysts [25,26]. It is mainly attributed to the following excellent properties of Ti3C2: (i) The Ti3C2 surface of MXenes, adorned with various terminal functional groups such as -OH, -O, and -F, can establish a strong chemical interface with semiconductors [27]. This interaction significantly contributes to inhibiting the recombination of electrons and holes [17]. (ii) The notable metallic conductivity of Ti3C2 ensures rapid carrier migration, thereby facilitating efficient separation of electrons and holes [28,29]. (iii) The exposed terminal metal sites on Ti3C2 are likely to result in higher reactivity compared to that of carbon-based materials [26,30]. Given the aforementioned characteristics, Ti3C2 has emerged as a favored co-catalyst in the synthesis of photocatalysts [31]. Li et al. reported that mesoporous TiO2 nanoparticles were successfully anchored onto a highly conductive Ti3C2 MXene co-catalyst using an electrostatic self-assembly strategy. The composite material, TiO2/Ti3C2, has demonstrated superior photocatalytic performance. This is particularly evident in its ability to degrade methyl orange, where it achieved an exceptionally high efficiency of 99.6%. Additionally, the material exhibits a significant hydrogen production rate, quantified at 218.85 μmol·g−1·h−1. These findings underscore the composite’s potential efficacy in photocatalytic applications [32]. Ran et al. synthesized Ti3C2 nanoparticles with CdS to induce a photocatalytic hydrogen production activity of 14,342 μmol·g−1·h−1. The enhanced performance observed can be attributed to the advantageous characteristics of the Ti3C2 nanoparticles. Specifically, favorable positioning of the Fermi level, coupled with the nanoparticles’ excellent electrical conductivity, play critical roles in facilitating the high efficiency of the composite. These properties significantly contribute to the overall superior photocatalytic performance of the material [28].
In this work, we synthesized a Bi2WO6 photoelectrode with a large proportional exposure of {001} facets on an ITO substrate in situ (Bi2WO6 NAs). Subsequently, the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode was synthesized via electrochemical deposition, layering NiFe-LDH and delaminating Ti3C2 onto Bi2WO6 NAs. Delaminated MXenes were utilized due to their superior electrical conductivity and robust mechanical properties, which are essential for the photocatalytic applications investigated. The layered structure facilitates efficient electron transport and provides numerous active sites for photocatalytic reactions, thereby enhancing the overall performance of the photocatalysts. This approach ensured precise material integration for the composite electrode. Taking the degradation of bisphenol A (BPA) as an example, the effect of the combination of Ti3C2 and NiFe-LDH on the catalytic performance of the Bi2WO6 NAs photoelectrode was investigated. Under PC, EC, and PEC conditions, the degradation activity of the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode is better than that of the Bi2WO6 NAs photoelectrode, and the order of activity from high to low is NiFe-LDH/Ti3C2/Bi2WO6 NAs > NiFe-LDH/Bi2WO6 NAs > Ti3C2/Bi2WO6 NAs > Bi2WO6 NAs. The extensive exposure of {001} facets in the Bi2WO6 nanoplate arrays (NAs) photoelectrode is advantageous for photocatalytic oxidation reactions. Moreover, the synergistic interaction between NiFe–layered double hydroxide (LDH) and Ti3C2 significantly enhances the separation efficiency of photogenerated carriers in Bi2WO6. Performance of the NiFe-LDH/Bi2WO6 photoelectrode and the mechanism of enhancing its PEC activity have been systematically studied.

2. Results and Discussion

2.1. Catalyst Characterization

The crystallographic structure and compositional analysis of the synthesized samples were examined utilizing X-ray diffraction (XRD) techniques. (Figure 1a). The XRD pattern shows three diffraction peaks at 8.9°, 18.3°, and 60.8°, corresponding to the (002), (004), and (110) planes of Ti3C2, respectively [29]. In addition, all of the diffraction peaks match well with the Bi2WO6 NAs according to JCPDS 39-0256 [33]. The observed intensity ratio of the {002} and {131} peaks in Bi2WO6 nanoplate arrays (NAs) is approximately 0.5, which is notably higher than the standard value of 0.19. This indicates that the as-prepared Bi2WO6 NAs exhibit anisotropic growth along the {001} planes. The extensive exposure of {001}-oriented Bi2WO6 NAs is advantageous, as it facilitates photocatalytic oxidation reactions [34]. After adding Ti3C2 (MXene) in the formation of Bi2WO6 NAs, the area integral ratio of diffraction peak between {002} and {131} crystal planes increases to 0.54, indicating that the existence of MXene expanded the {001}-orientated crystal faces exposure of Bi2WO6 NAs. The observed negativity of the Ti3C2 surface can be attributed to the adsorption of a substantial quantity of -F and -OH groups. This characteristic is particularly relevant during the preparation of MXene, where Bi3+ ions are adsorbed onto the Ti3C2 surface, further influencing its chemical behavior and interactions. Therefore, Bi2WO6 NAs expose more [Bi2O2]2+ layers parallel to the {001} crystal plane. In addition, the diffraction peak of Bi2WO6 NAs becomes wider, which is due to the growth-limiting effect of MXene on Bi2WO6 NAs [17]. UV-Vis Diffuse Reflectance Spectroscopy (DRS) spectra provide insights into the optical properties of the as-prepared samples. As shown in Figure 1b, Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrodes have obvious absorption responses in the visible light region. Moreover, the absorption edge of Ti3C2/Bi2WO6 NAs and NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrodes is near 450 nm, indicating that the combination of Ti3C2 broadens the response of Bi2WO6 NAs to visible light.
Scanning electron microscopy (SEM) images (Figure 2a and Figure S1) exhibit that Bi2WO6 nanosheets are vertically grown on the ITO substrate with a size of about 200 nm. The nanoplates are interconnected, forming a framework array structure. Following compounding with Ti3C2, the size of Bi2WO6 nanosheets is observed to be approximately 100 nm. This dimension is attributed to the growth restriction imposed by Ti3C2 on Bi2WO6. (Figure 2b,c) [17]. Figure 2d displays the structural morphology of the NiFe–layered double hydroxide (LDH)/Ti3C2/Bi2WO6 nanoplate arrays (NAs) photoelectrode. NiFe-LDH is observed to be uniformly coated on the Ti3C2/Bi2WO6 surface, presenting a flocculent morphology that is evenly dispersed. As shown in Supplementary Figure S2, the uniform detection of elements such as tungsten (W), bismuth (Bi), oxygen (O), nickel (Ni), iron (Fe), titanium (Ti), and carbon (C) confirms the even distribution of NiFe-LDH and Ti3C2 across the Bi2WO6 NAs.

2.2. Enhanced PEC Degradation Activity

Catalytic performance of as-prepared catalysts is estimated by the degradation of BPA in aqueous solution under PC, EC, and PEC systems. As can be seen from Figure S3, BPA does not undergo self-degradation under visible light irradiation. The EC activity of the Bi2WO6 NAs photoelectrode also shows a low degradation rate of BPA (Figure 3a and Figure 4a). Generally speaking, Ti3C2 and NiFe-LDH are excellent electrocatalytic materials with high conductivity, so the EC activity is significantly improved after compounding with Ti3C2 and NiFe-LDH. Upon exposure exclusively to visible light, the degradation rate constant, denoted as k, of the NiFe-LDH/Ti3C2/Bi2WO6 nanocomposite photoelectrode exhibits an enhancement by a factor of approximately 10 compared to the degradation rate constant of the pristine Bi2WO6 nanocomposite photoelectrode (Figure 3b and Figure 4b) due to the excellent electron transport properties of Ti3C2 and the fast trapping of photogenerated holes by NiFe-LDH. In the PEC system, the catalytic performance of Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrodes is further improved, all of which are higher than the PC or EC system alone (Figure 3c and Figure 4c). The external potential can quickly transfer the photogenerated electrons generated by Bi2WO6 NAs to the counter electrode through the external circuit. Consequently, PEC activity escalates with the augmentation of external potential. At external potentials below the water decomposition threshold, such potential facilitates the migration of electrons towards the counter electrode, thereby enhancing the segregation of photogenerated electrons and holes. Conversely, when the external potential surpasses the threshold for water decomposition, it instigates complex oxidation processes. Figure S4 shows the PEC performance of the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode under different external potentials. Under the irradiation of visible light (λ ≥ 420 nm), the PEC degradation rate of BPA gradually increases with increasing external potential. However, when the external potential reaches 3.0 V, the surface of the Bi2WO6 NAs photoelectrode blackens rapidly, leading to deactivation of the catalyst and a sharp decrease in degradation activity.

2.3. Enhancement Mechanism of Photoelectrochemical (PEC) Activity

To elucidate the underlying mechanism behind the augmented PEC activity of NiFe-LDH/Ti3C2/Bi2WO6 nanocomposites, an array of photoelectrical characterizations has been conducted. As can be seen from Figure 5a, the photocurrent density of the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode reaches about 3.0 μA/cm2, which represents an enhancement of approximately 3.8-fold compared to the Ti3C2/Bi2WO6 nanocomposite photoelectrode and a 6.0-fold increase over the pristine Bi2WO6 nanocomposite photoelectrode. The increase of photocurrent density of the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode is attributed to the synergistic effect of NiFe-LDH and Ti3C2, which greatly improves the separation efficiency of Bi2WO6 NAs photogenerated charge. Electrochemical impedance spectroscopy is utilized to investigate the impedance characteristics of the synthesized samples, as depicted in Figure 5b. The curve radius of the pure Bi2WO6 NAs photoelectrode is the largest, and the impedance value is also the largest. After compounding with Ti3C2 and NiFe-LDH, the curve radius of the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode decreases and the impedance value decreases accordingly, which is consistent with the above transient photocurrent response results. The catalytic performance of the as-prepared sample is investigated by linear sweep voltammetry (LSV) (Figure 5c). The order of current density under different external potentials is NiFe-LDH/Ti3C2/Bi2WO6 NAs > NiFe-LDH/Bi2WO6 NAs > Ti3C2/Bi2WO6 NAs > Bi2WO6 NAs. Furthermore, the starting potential of the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode has an obvious negative shift, indicating that the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode has a stronger redox ability. From Figure 5d, the curve slopes of the Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrodes are all positive, which demonstrates that Bi2WO6 is an N-type semiconductor. At the same time, after compounding with Ti3C2 and NiFe-LDH, the flat band potential shifts significantly, illustrating that the band edge bending of the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode decreases. It is beneficial to promote charge transfer at the electrode/electrolyte interface, accelerate the surface reaction kinetics process, and then enhance catalytic performance. To investigate the correlation between the photoelectric properties and light absorption capabilities of the synthesized samples, the photoelectric conversion efficiency of Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrodes are calculated (Figure S5). All samples have a photoelectric conversion ability in the ultraviolet region, and the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode has the strongest photoelectric conversion ability.
Figure 6 presents the outcomes of X-ray photoelectron spectroscopy (XPS) analyses performed on the catalysts. It was found that Bi, W, O, Ni, Fe, Ti, and C elements are present in the NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrode (Figure 6a). Figure 6b illustrates the predominance of Ti-C bonds within the Ti2p peaks, alongside a minor presence of Ti-O structures. This suggests that while there may be slight oxidation of the Ti3C2 material, it does not detract from the dominant influence of the Ti3C2 structure in the ternary catalytic system. Figure 6c,d show the chemical states of Bi 4f and W 4f of as-prepared samples. Compared with that of Bi2WO6 NAs, the binding energy of Ti3C2/Bi2WO6 NAs in Bi 4f and W 4f spectra show positive shifts. However, the peak shifts of Bi 4f and Bi 4f5/2 in the NiFe-LDH/Bi2WO6 NAs photoelectrode are opposite [13,35]. Moreover, the O 1s spectra of the as-prepared sample in Figure 6e consists of Bi-O, W-O, and surface adsorbed oxygen [36]. The peak positions of Bi-O and W-O in the Ti3C2/Bi2WO6 NAs photoelectrode show positive shifts, and the position of adsorbed oxygen shifts to 531.78 eV. On the contrary, the peak positions of Bi-O and W-O in the NiFe-LDH/Bi2WO6 NAs photoelectrode shift negatively to 529.22 eV and 530.04 eV, and the peaks at 530.55 eV, 531.72 eV, and 532.30 eV belong to surface adsorbed oxygen, interlayer anion (OH), and H2O of NiFe-LDH, respectively. The above proves that Bi2WO6 NAs interact with Ti3C2 and NiFe-LDH, respectively, inducing the migration of photogenerated electrons from Bi2WO6 NAs to Ti3C2 and photogenerated holes from Bi2WO6 NAs to NiFe-LDH. Furthermore, an upward shift in the peak positions of Ni 2p and Fe 2p observed in NiFe-LDH/Bi2WO6 nanocomposite photoelectrodes and NiFe-LDH/Ti3C2/Bi2WO6 nanocomposite photoelectrodes (as shown in Figure 6f,g) corroborates the interaction between the surface energy of Bi2WO6 nanocomposites and NiFe-LDH. This charge transfer mechanism aids in the efficient separation of electrons and holes, thereby improving photoelectrochemical performance.
Building on the preceding discussion, a more detailed mechanism is proposed for the photoelectrochemical (PEC) degradation of organic pollutants using the NiFe-LDH/Ti3C2/Bi2WO6 nanocomposite photoelectrode, as illustrated in Figure 7. The ternary composite system comprising Ti3C2, Bi2WO4, and layered double hydroxide (LDH) forms a unique and sophisticated heterojunction that exhibits characteristics of both Type-II and Z-scheme heterojunctions. This innovative configuration leverages the distinct electronic properties of each component to facilitate efficient charge separation and transfer, thereby enhancing photocatalytic performance under light irradiation. The Ti3C2/Bi2WO4 interface forms a Type-II heterojunction, where alignment of their conduction and valence bands allows for the spatial separation of photogenerated electrons and holes. Electrons tend to migrate towards Ti3C2, while holes accumulate in Bi2WO4, thus reducing the recombination rate and enhancing photocatalytic efficiency. The addition of LDH into the Ti3C2/Bi2WO4 system introduces a Z-scheme mechanism, particularly when LDH acts as a bridge for electron transfer between Ti3C2 and Bi2WO4. This configuration preserves the high reduction potential of Ti3C2’s electrons and the high oxidation potential of Bi2WO4’s holes, making the composite highly effective for redox reactions.
The LDH/Ti3C2/Bi2WO6 ternary composite system leverages a unique combination of Type-II and Z-scheme heterojunctions, offering significant advantages for photocatalytic applications. This sophisticated heterostructure ensures enhanced charge separation and transfer, effectively minimizing recombination and maximizing the availability of reactive charge carriers for photocatalysis. Additionally, the incorporation of these materials broadens the light absorption range, enabling the composite to utilize a larger portion of the solar spectrum, particularly under visible light, which significantly improves its photocatalytic activity. The Z-scheme configuration within this ternary system preserves the high reduction and oxidation potentials of electrons and holes, respectively, enabling the composite to efficiently participate in a wide range of redox reactions, including the degradation of pollutants and water splitting. Moreover, the synergistic effects among Ti3C2, Bi2WO4, and LDH not only enhance the composite’s stability and durability under photocatalytic conditions but also ensure sustained activity over extended periods, making the LDH/Ti3C2/Bi2WO6 system a highly efficient and versatile option for environmental remediation and energy conversion technologies.
The enhancement of PEC activity in this nanocomposite photoelectrode is the result of multiple synergistic effects, notably the following: (i) The enhanced exposure of {001} facets in Bi2WO6 nanocomposites within the NiFe-LDH/Ti3C2/Bi2WO6 framework significantly contributes to an increased generation of holes. This feature not only facilitates more active participation in the degradation process but also substantially strengthens photocatalytic oxidation reactions. The {001} facets of Bi2WO6 are known for their high photocatalytic activity, and their prominent exposure within the composite structure maximizes the photocatalytic sites available for reaction; (ii) The strategic integration of NiFe-LDH with Ti3C2 creates a highly effective conduit for the swift separation and transfer of photogenerated charges. Under illumination, the electrons excited from the conduction band of Bi2WO6 are quickly transferred to Ti3C2. This transition is facilitated by the close interfacing and strong electronic interaction between Bi2WO6 and Ti3C2, which serves as an effective conduit for the rapid transportation of electrons to the counter electrode via the external circuit. This bridging significantly curtails the recombination of photogenerated electron–hole pairs, thereby enhancing the overall separation efficiency of charge carriers; (iii) The directed migration of photogenerated holes towards the NiFe-LDH surface plays a crucial role in the oxidative degradation of organic pollutants. NiFe-LDH acts as a catalyst that accelerates the oxidation process, converting organic contaminants into carbon dioxide (CO2) and water (H2O). This process is facilitated by the high affinity of NiFe-LDH for photogenerated holes, which promotes their accumulation on the LDH surface, thereby increasing local hole concentration and enhancing the oxidation reaction rate. As a result, the NiFe-LDH/Ti3C2/Bi2WO6 nanocomposite photoelectrode represents a sophisticated integration of materials that leverages the unique properties of each component to maximize PEC degradation efficiency. The optimized exposure of photocatalytically active facets, coupled with the strategic facilitation of charge separation and targeted reaction pathways, underpin the superior performance of this ternary composite in the PEC degradation of organic pollutants. This comprehensive understanding of the degradation mechanism provides valuable insights into designing more efficient PEC systems for environmental remediation.

3. Conclusions

In summary, the NiFe-LDH/Ti3C2/Bi2WO6 nanocomposite photoelectrode, featuring a significant proportion of {001}-oriented facets, was successfully synthesized through a solvothermal reaction followed by electrochemical deposition. Comparative analysis under electrochemical (EC), photocatalytic (PC), and photoelectrochemical (PEC) conditions revealed that the NiFe-LDH/Ti3C2/Bi2WO6 nanocomposite photoelectrode exhibits superior degradation performance relative to the pristine Bi2WO6 nanocomposite photoelectrode. The activity ranking, in descending order, is as follows: NiFe-LDH/Ti3C2/Bi2WO6 NAs > NiFe-LDH/Bi2WO6 NAs > Ti3C2/Bi2WO6 NAs > Bi2WO6 NAs. The augmented photocatalytic efficacy of the NiFe-LDH/Ti3C2/Bi2WO6 nanocomposite photoelectrode can be ascribed to the synergistic interplay between Ti3C2 and NiFe-LDH, which significantly enhances the charge separation efficiency and reaction kinetics of Bi2WO6 NAs. This study elucidates a novel approach for enhancing reaction kinetics in the PEC degradation of pollutants.

4. Materials and Methods

4.1. Preparation of Bi2WO6 NAs Photoelectrode

Initially, 0.9 g of polyvinylpyrrolidone (PVP) was dissolved in 30 mL of ethylene glycol at a temperature of 95 °C in two separate beakers to achieve a homogeneous, transparent solution. Subsequently, 0.3299 g of sodium tungstate dihydrate (Na2WO4·2H2O) was dissolved in the first beaker (hereafter referred to as solution 1), while 0.9701 g of bismuth nitrate pentahydrate (Bi(NO3)3·5H2O) was dissolved in the second beaker (referred to as solution 2). Solution 1 was then gradually introduced into solution 2 with constant stirring to ensure thorough mixing. The resultant mixture was subjected to a pre-reaction phase at 98 °C for 1.5 h with sustained stirring. Following this, the mixture was transferred into a Teflon-lined autoclave, in which cleaned indium tin oxide (ITO) substrates had been vertically positioned, and subsequently heated at 178 °C for 24 h. Upon cooling to ambient temperature, the ITO substrates were carefully removed, and washed multiple times with ethanol and deionized water to obtain the Bi2WO6 nanoplate arrays (NAs) photoelectrode.

4.2. Preparation of Ti3C2

A mass of 1.0 g of MAX phase powder (Ti3AlC2) was dispersed in a Teflon container. To this, 10 mL of 40% HF solution was added per gram of MAX phase, ensuring the mixture was stirred gently to achieve uniform etching. The reaction was allowed to proceed for 24 h at room temperature under constant stirring to facilitate the complete removal of the A layer. Following etching, the mixture was diluted with deionized water and allowed to settle, enabling the separation of undissolved solids. The supernatant, containing excess HF and reaction byproducts, was carefully decanted. This washing process was repeated several times until the pH of the supernatant reached neutrality, indicating the effective removal of residual HF. The resulting MXene was then collected by centrifugation at 3500 rpm for 5 min, washed with deionized water, and spread out for drying at room temperature in a fume hood or under vacuum at 60 °C. The dried MXene was stored in an airtight container under N2 atmosphere to prevent oxidation.

4.3. Preparation of NiFe-LDH

A volume of 70 mL of deionized water was introduced into an electrolytic cell, followed by the addition of 0.15 mol of nickel nitrate hexahydrate (Ni(NO3)2·6H2O) and 0.15 mol of iron(II) sulfate heptahydrate (FeSO4·7H2O). An indium tin oxide (ITO) substrate, a platinum wire, and an Ag/AgCl electrode were utilized as the working electrode, counter electrode, and reference electrode, respectively. The electrochemical deposition was conducted at a voltage of −1.0 V for a duration of 300 s. Subsequent to the electrochemical deposition, the ITO substrate was removed and exposed to ambient air to facilitate the natural oxidation of Fe2+ ions to Fe3+ ions. This process culminated in the formation of the NiFe–layered double hydroxide (LDH) electrode.

4.4. Preparation of Ti3C2/Bi2WO6 NAs Photoelectrode

Initially, 20 mL of ethylene glycol was introduced into two beakers, and subsequently placed in a 95 °C water bath. Nitrogen gas was infused into each beaker to establish an inert environment. To these, 0.6 g of polyvinylpyrrolidone (PVP) was added as a dispersant. Specifically, beaker 2 received 0.0093 g of Ti3C2 powder, with nitrogen flow maintained to protect the sensitive material. The mixture was sealed and magnetically stirred for 20 min, ensuring uniform dispersion of Ti3C2.
Simultaneously, 0.3299 g of Na2WO4·2H2O and 0.9701 g of Bi(NO3)3·5H2O were added to beakers 1 and 2, respectively, under nitrogen to prevent any oxidative reactions. The contents of beaker 1 were slowly added to beaker 2, and the combined solution was pre-reacted for one hour at 95 °C, still under nitrogen. This mixture was then transferred to a 50 mL autoclave containing a pre-treated indium tin oxide (ITO) substrate, with the entire assembly purged with nitrogen. The autoclave underwent hydrothermal treatment at 180 °C for 24 h, facilitating formation of the composite on the ITO, all while under nitrogen to minimize degradation. Upon cooling to room temperature under nitrogen, the ITO substrate, coated with the Ti3C2/Bi2WO6 composite, was meticulously rinsed with deionized water and anhydrous ethanol, and dried in a vacuum oven at 60 °C.

4.5. Preparation of NiFe-LDH/Bi2WO6 NAs Photoelectrode

A volume of 70 mL of deionized water was added to an electrolytic cell. Then, under a nitrogen atmosphere and with continuous stirring, 0.15 M of Ni(NO3)2·6H2O and 0.15 M of FeSO4·7H2O were added, respectively. The flow of N2 gas was to prevent the oxidation of Fe2+. The electrochemical deposition of NiFe-LDH was carried out using a three-electrode system. A Bi2WO6 photoelectrode, a platinum wire, and an Ag/AgCl electrode were used as the working electrode, counter electrode, and reference electrode, respectively. The deposition voltage was set to −1.0 V, and the deposition time was 30 s. After the electrochemical deposition finished, the ITO was removed, rinsed with deionized water, and then placed in air to allow Fe2+ to naturally oxidize to Fe3+. Finally, a NiFe-LDH/Bi2WO6 or NiFe-LDH/Ti3C2/Bi2WO6 composite photoelectrode was obtained.

4.6. Preparation of NiFe-LDH/Ti3C2/Bi2WO6 NAs Photoelectrode

A Ti3C2/Bi2WO6 NAs photoelectrode was used as the working electrode, and the experimental steps were repeated as in Section 4.5.

4.7. Characterization Techniques

X-ray diffraction (XRD) analyses were conducted using a Bruker D8-Focus diffractometer equipped with Cu Kα radiation. The morphology of the synthesized sample was examined via a scanning electron microscope (SEM, SU8010, Hitachi Ltd., Chiyoda, Japan) across an acceleration voltage range of 200 V to 30 kV, supplemented by energy dispersive X-ray (EDX) spectroscopy for elemental analysis. The interplanar spacing of the lattice fringes was determined using a high-resolution transmission electron microscope (HRTEM, Tecnai F20, FEI Company, Hillsboro, OR, USA). Ultraviolet-visible diffuse reflectance spectra (DRS) were acquired on a Hitachi 4100 spectrophotometer, employing BaSO4 as the reference standard. X-ray photoelectron spectroscopy (XPS) measurements were performed on a PHI Quantera ULVAC XPS system (ULVAC, Inc., Chigasaki, Japan). Photoelectrocatalytic (PEC) characterizations were conducted in a quartz electrochemical cell, incorporating a three-electrode setup, with an electrochemical workstation (CHI 660D, CH Instruments, Inc., Shanghai, China). A 0.1 M solution of Na2SO4 served as the electrolyte. Electrochemical impedance spectroscopy (EIS) was executed over a frequency range from 0.005 Hz to 105 Hz, applying a sinusoidal AC disturbance signal of 5 mV to probe the electrochemical properties of the materials.

4.8. Degradation Activity Test

The PC and PEC activities of the synthesized samples were assessed through the degradation of bisphenol A (BPA) at a concentration of 10 mol·L−1 in a 50 mL solution of Na2SO4 (0.1 mol·L−1). Prior to initiating the PC and PEC reactions, the solution was stirred in darkness for 30 min to establish adsorption–desorption equilibrium. For the PEC activity evaluation, a photoelectrode, a Pt wire, and a saturated calomel electrode (SCE) were employed as the working electrode, counter electrode, and reference electrode, respectively. The photoelectrode was exposed to either simulated sunlight or visible light (λ > 420 nm), generated by a 300 W Xe lamp (PLS-SXE300C/300CUV, Perfect Light Ltd, Beijing, China), with an average intensity of 100 mW cm−2. For comparative analysis, the samples were also subjected to electrochemical (EC) degradation of BPA under identical conditions but without light irradiation. The degradation process was conducted over a duration of 4 h, with aliquots of 2.5 mL sampled every 30 min. The concentration of BPA was determined using a high-performance liquid chromatography (HPLC) system (Shimadzu LC-20AT, Shimadzu Corporation, Kyoto, Japan).

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/nano14050477/s1: Figure S1. SEM image of Bi2WO6 NAs; Figure S2. EDX mapping of NiFe-LDH/Ti3C2/Bi2WO6 NAs; Figure S3. Comparison of EC, PC, and PEC degradation rates over Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs (λ ≥ 420 nm, external potential = 1.0 V); Figure S4. Comparison of the PEC degradation rate of BPA over NiFe-LDH/Ti3C2/Bi2WO6 NAs at different bias voltages (λ ≥ 420 nm); Figure S5. Photoelectric conversion efficiency of Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs.

Author Contributions

Conceptualization, L.L. and J.Z.; methodology, R.L.; software, R.L..; validation, Y.W. (Yuting Wang) and R.L.; formal analysis, Y.W. (Yuting Wang); investigation, Y.W. (Yuting Wang) and R.L.; resources, W.H.; data curation, R.L. and Y.W. (Yuting Wang); writing—original draft preparation, Y.W. (Yuting Wang) and R.L.; writing—review and editing, Y.W. (Yuting Wang); visualization, Y.W. (Yuting Wang); supervision, Y.W. (Yajun Wang); project administration, Y.W. (Yajun Wang); funding acquisition, Y.W. (Yajun Wang) and Y.W. (Yuting Wang). All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Key Research and Development Program of China (2021YFB4000405, 2019YFC1904500), National Natural Science Foundation of China (52270115, 52236003, 52203132, 21777080)

Data Availability Statement

Data supporting the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fan, G.; Yang, S.; Du, B.; Luo, J.; Lin, X.; Li, X. Sono-photo hybrid process for the synergistic degradation of levofloxacin by FeVO4/BiVO4: Mechanisms and kinetics. Environ. Res. 2022, 204, 112032. [Google Scholar] [CrossRef]
  2. Shen, R.; He, K.; Zhang, A.; Li, N.; Ng, Y.H.; Zhang, P.; Hu, J.; Li, X. In-situ construction of metallic Ni3C@Ni core-shell cocatalysts over g-C3N4 nanosheets for shell-thickness-dependent photocatalytic H2 production. Appl. Catal. B-Environ. 2021, 291, 120104. [Google Scholar] [CrossRef]
  3. Fei, T.; Yu, L.; Liu, Z.; Song, Y.; Xu, F.; Mo, Z.; Liu, C.; Deng, J.; Ji, H.; Cheng, M.; et al. Graphene quantum dots modified flower like Bi2WO6 for enhanced photocatalytic nitrogen fixation. J. Colloid Interface Sci. 2019, 557, 498–505. [Google Scholar] [CrossRef]
  4. Zhang, Z.; Wang, W.; Yin, W.; Shang, M.; Wang, L.; Sun, S. Inducing photocatalysis by visible light beyond the absorption edge: Effect of upconversion agent on the photocatalytic activity of Bi2WO6. Appl. Catal. B-Environ. 2010, 101, 68–73. [Google Scholar] [CrossRef]
  5. Ren, G.M.; Han, H.T.; Wang, Y.X.; Liu, S.T.; Zhao, J.Y.; Meng, X.C.; Li, Z.Z. Recent Advances of Photocatalytic Application in Water Treatment: A Review. Nanomaterials 2021, 11, 1804. [Google Scholar] [CrossRef]
  6. Bessy, T.C.; Bindhu, M.R.; Johnson, J.; Chen, S.-M.; Chen, T.-W.; Almaary, K.S. UV light assisted photocatalytic degradation of textile waste water by Mg0.8-xZnxFe2O4 synthesized by combustion method and in-vitro antimicrobial activities. Environ. Res. 2022, 204, 111917. [Google Scholar] [CrossRef]
  7. Wu, S.J.; Sun, J.G.; Li, Q.; Hood, Z.D.; Yang, S.; Su, T.M.; Peng, R.; Wu, Z.; Sun, W.W.; Kent, P.R.C.; et al. Effects of Surface Terminations of 2D Bi2WO6 on Photocatalytic Hydrogen Evolution from Water Splitting. Acs Appl. Mater. Interfaces 2020, 12, 20067–20074. [Google Scholar] [CrossRef] [PubMed]
  8. Kaur, A.; Sharma, S.; Sood, S.; Umar, A.; Bhinder, S.S.; Kansan, S.K. Visible Light Driven Photo-Catalytic Degradation of Fluoroquinolone Antibiotic Drug Using Bi2WO6 Spheres Composed of Fluffy Nanosheets. Nanosci. Nanotechnol. Lett. 2016, 8, 660–666. [Google Scholar] [CrossRef]
  9. Qi, S.; Zhang, R.; Zhang, Y.; Liu, X.; Xu, H. Preparation and photocatalytic properties of Bi2WO6/g-C3N4. Inorg. Chem. Commun. 2021, 132, 108761. [Google Scholar] [CrossRef]
  10. Nguyen Dang, P.; Luc Huy, H.; Guo, P.-C.; Chen, X.-B.; Chou, W.C. Study of photocatalytic activities of Bi2WO6/BiVO4 nanocomposites. J. Sol-Gel Sci. Technol. 2017, 83, 640–646. [Google Scholar] [CrossRef]
  11. Bai, J.; Ren, X.; Chen, X.; Lu, P.; Fu, M. Oxygen Vacancy-Enhanced Ultrathin Bi2O3-Bi2WO6 Nanosheets’ Photocatalytic Performances under Visible Light Irradiation. Langmuir 2021, 37, 5049–5058. [Google Scholar] [CrossRef]
  12. Li, Z.S.; Luo, M.H.; Li, B.L.; Lin, Q.C.; Liao, X.C.; Yu, H.Q.; Yu, C.L. 3-D hierarchical micro/nano-structures of porous Bi2WO6: Controlled hydrothermal synthesis and enhanced photocatalytic performances. Microporous Mesoporous Mater. 2021, 313, 110830. [Google Scholar] [CrossRef]
  13. Cao, P.F.; Ma, S.Y.; Xu, X.L. Novel ultra-sensitive dandelion-like Bi2WO6 nanostructures for ethylene glycol sensing application. Vacuum 2020, 181, 109748. [Google Scholar] [CrossRef]
  14. Lu, Y.; Zhao, K.; Zhao, Y.H.; Zhu, S.Y.; Yuan, X.; Huo, M.X.; Zhang, Y.; Qiu, Y. Bi2WO6/TiO2/Pt nanojunction system: A UV-vis light responsive photocatalyst with high photocatalytic performance. Colloids Surf. A-Physicochem. Eng. Asp. 2015, 481, 252–260. [Google Scholar] [CrossRef]
  15. Chen, X.; Chen, C.; Zang, J. Bi2MoO6 nanoflower-like microsphere photocatalyst modified by boron-doped carbon quantum dots: Improving the photocatalytic degradation performance of BPA in all directions. J. Alloys Compd. 2023, 962, 171167. [Google Scholar] [CrossRef]
  16. Zhang, C.; Zhu, Y.F. Synthesis of square Bi2WO6 nanoplates as high-activity visible-light-driven photocatalysts. Chem. Mater. 2005, 17, 3537–3545. [Google Scholar] [CrossRef]
  17. Cao, S.W.; Shen, B.J.; Tong, T.; Fu, J.W.; Yu, J.G. 2D/2D Heterojunction of Ultrathin MXene/Bi2WO6 Nanosheets for Improved Photocatalytic CO2 Reduction. Adv. Funct. Mater. 2018, 28, 1800136. [Google Scholar] [CrossRef]
  18. Ali, B.; Abdelkader, E.; Naceur, B.; Houcined, C.; Nadjia, L.; Nourredine, B. Sunlight-driven photocatalytic degradation of Rhodamine B by BiOCl and TiO2 deposited on NiCr-LDH. Int. J. Environ. Anal. Chem. 2021, 103, 6722–6741. [Google Scholar] [CrossRef]
  19. He, Y.; Zhou, S.; Wang, Y.; Jiang, G.; Jiao, F. Fabrication of g-C3N4@NiFe-LDH heterostructured nanocomposites for highly efficient photocatalytic removal of rhodamine B. J. Mater. Sci.-Mater. Electron. 2021, 32, 21880–21896. [Google Scholar] [CrossRef]
  20. Suppaso, C.; Pongkan, N.; Intachai, S.; Ogawa, M.; Khaorapapong, N. Magnetically recoverable beta-Ni(OH)2/gamma-Fe2O3/NiFe-LDH composites; isotherm, thermodynamic and kinetic studies of synthetic dye adsorption and photocatalytic activity. Appl. Clay Sci. 2021, 213, 106115. [Google Scholar] [CrossRef]
  21. Wang, R.; Su, S.; Ren, X.; Guo, W. Polyoxometalate intercalated La-doped NiFe-LDH for efficient removal of tetracycline via peroxymonosulfate activation. Sep. Purif. Technol. 2021, 274, 119113. [Google Scholar] [CrossRef]
  22. Zhang, D.; Dong, W.; Liu, Y.; Gu, X.; Yang, T.; Hong, Q.; Li, D.; Zhang, D.; Zhou, H.; Huang, H.; et al. Ag-In-Zn-S Quantum Dot-Dominated Interface Kinetics in Ag-In-Zn-S/NiFe LDH Composites toward Efficient Photoassisted Electrocatalytic Water Splitting. Acs Appl. Mater. Interfaces 2021, 13, 42125–42137. [Google Scholar] [CrossRef] [PubMed]
  23. Zhu, Y.; Ren, J.; Yang, X.; Chang, G.; Bu, Y.; Wei, G.; Han, W.; Yang, D. Interface engineering of 3D BiVO4/Fe-based layered double hydroxide core/shell nanostructures for boosting photoelectrochemical water oxidation. J. Mater. Chem. A 2017, 5, 9952–9959. [Google Scholar] [CrossRef]
  24. Aboubakr, A.E.A.; El Rouby, W.M.A.; Khan, M.D.; Farghali, A.A.; Revaprasadu, N. ZnCr-CO3 LDH/ruptured tubular g-C3N4 composite with increased specific surface area for enhanced photoelectrochemical water splitting. Appl. Surf. Sci. 2020, 508, 145100. [Google Scholar] [CrossRef]
  25. Chen, S.; Liu, R.; Kuai, Z.; Li, X.; Lian, S.; Jiang, D.; Tang, J.; Li, L.; Wu, R.; Peng, C. Facile synthesis of a novel BaSnO3/MXene nanocomposite by electrostatic self-assembly for efficient photodegradation of 4-nitrophenol. Environ. Res. 2022, 204, 111949. [Google Scholar] [CrossRef] [PubMed]
  26. Zhao, W.J.; Qin, J.Z.; Yin, Z.F.; Hu, X.; Liu, B.J. 2D MXenes for Photocatalysis. Prog. Chem. 2019, 31, 1729–1736. [Google Scholar]
  27. Adomavičiūtė-Grabusovė, S.; Ramanavičius, S.; Popov, A.; Šablinskas, V.; Gogotsi, O.; Ramanavičius, A. Selective Enhancement of SERS Spectral Bands of Salicylic Acid Adsorbate on 2D Ti3C2Tx-Based MXene Film. Chemosensors 2021, 9, 223. [Google Scholar] [CrossRef]
  28. Ran, J.; Gao, G.; Li, F.-T.; Ma, T.-Y.; Du, A.; Qiao, S.-Z. Ti3C2 MXene co-catalyst on metal sulfide photo-absorbers for enhanced visible-light photocatalytic hydrogen production. Nat. Commun. 2017, 8, 13907. [Google Scholar] [CrossRef]
  29. Pasupuleti, K.S.; Thomas, A.M.; Vidyasagar, D.; Rao, V.N.; Yoon, S.G.; Kim, Y.-H.; Kim, S.-G.; Kim, M.-D. ZnO@Ti3C2Tx MXene Hybrid Composite-Based Schottky-Barrier-Coated SAW Sensor for Effective Detection of Sub-ppb-Level NH3 at Room Temperature under UV Illumination. ACS Mater. Lett. 2023, 5, 2739–2746. [Google Scholar] [CrossRef]
  30. Mahar, I.; Memon, F.H.; Lee, J.-W.; Kim, K.H.; Ahmed, R.; Soomro, F.; Rehman, F.; Memon, A.A.; Thebo, K.H.; Choi, K.H. Two-Dimensional Transition Metal Carbides and Nitrides (MXenes) for Water Purification and Antibacterial Applications. Membranes 2021, 11, 869. [Google Scholar] [CrossRef]
  31. Huang, K.L.; Li, C.H.; Li, H.Z.; Ren, G.M.; Wang, L.; Wang, W.T.; Meng, X.C. Photocatalytic Applications of Two-Dimensional Ti3C2 MXenes: A Review. Acs Appl. Nano Mater. 2020, 3, 9581–9603. [Google Scholar] [CrossRef]
  32. Huang, Z.; Zhai, H.; Li, M.; Liu, X. Effects of Ti3AlC2 Content on Cu/Ti3AlC2 Composites by Hot Pressing Method. Rare Met. Mater. Eng. 2011, 40, 529–532. [Google Scholar]
  33. Zhang, G.; Hu, Z.; Sun, M.; Liu, Y.; Liu, L.; Liu, H.; Huang, C.-P.; Qu, J.; Li, J. Formation of Bi2WO6 Bipyramids with Vacancy Pairs for Enhanced Solar-Driven Photoactivity. Adv. Funct. Mater. 2015, 25, 3726–3734. [Google Scholar] [CrossRef]
  34. Yang, C.; Huang, Y.; Li, T.; Li, F. Bi2WO6 Nanosheets Synthesized by a Hydrothermal Method: Photocatalytic Activity Driven by Visible Light and the Superhydrophobic Property with Water Adhesion. Eur. J. Inorg. Chem. 2015, 2015, 2560–2564. [Google Scholar] [CrossRef]
  35. Song, X.C.; Zhou, H.; Huang, W.Z.; Wang, L.; Zheng, Y.F. Enhanced Photocatalytic Activity of Nano-Bi2WO6 by Tin Doping. Curr. Nanosci. 2015, 11, 627–632. [Google Scholar] [CrossRef]
  36. Lv, H.; Liu, Y.; Hu, J.; Li, Z.; Lu, Y. Ionic liquid-assisted hydrothermal synthesis of Bi2WO6-reduced graphene oxide composites with enhanced photocatalytic activity. Rsc Adv. 2014, 4, 63238–63245. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of ITO, NiFe-LDH, Ti3C2, Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs; (b) UV-Vis DRS spectra of NiFe-LDH, Ti3C2, Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs.
Figure 1. (a) XRD patterns of ITO, NiFe-LDH, Ti3C2, Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs; (b) UV-Vis DRS spectra of NiFe-LDH, Ti3C2, Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs.
Nanomaterials 14 00477 g001
Figure 2. SEM images of (a) Bi2WO6 NAs; (b,c) Ti3C2/Bi2WO6 NAs; (d) NiFe-LDH/Ti3C2/Bi2WO6 NAs.
Figure 2. SEM images of (a) Bi2WO6 NAs; (b,c) Ti3C2/Bi2WO6 NAs; (d) NiFe-LDH/Ti3C2/Bi2WO6 NAs.
Nanomaterials 14 00477 g002
Figure 3. Comparison of (a) EC, (b) PC, and (c) PEC degradation over time, and its first-order reaction fitting curves over Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrodes (λ ≥ 420 nm, external potential = 1.0 V). The associated rate constant, denoted as k.
Figure 3. Comparison of (a) EC, (b) PC, and (c) PEC degradation over time, and its first-order reaction fitting curves over Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrodes (λ ≥ 420 nm, external potential = 1.0 V). The associated rate constant, denoted as k.
Nanomaterials 14 00477 g003
Figure 4. Comparison of (a) EC, (b) PC, and (c) PEC degradation rates over Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrodes.
Figure 4. Comparison of (a) EC, (b) PC, and (c) PEC degradation rates over Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs photoelectrodes.
Nanomaterials 14 00477 g004
Figure 5. (a) Transient photocurrent density; (b) EIS Nyquist plots; (c) linear sweep voltammetry plots; (d) Mott–Schottky plots of Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs (λ ≥ 420 nm).
Figure 5. (a) Transient photocurrent density; (b) EIS Nyquist plots; (c) linear sweep voltammetry plots; (d) Mott–Schottky plots of Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs (λ ≥ 420 nm).
Nanomaterials 14 00477 g005
Figure 6. XPS of (a) survey spectra, (b) Ti 2p, (c) Bi 4f, (d) W 4f, (e) O 1s, (f) Ni 2p, and (g) Fe 2p of NiFe-LDH, Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs.
Figure 6. XPS of (a) survey spectra, (b) Ti 2p, (c) Bi 4f, (d) W 4f, (e) O 1s, (f) Ni 2p, and (g) Fe 2p of NiFe-LDH, Bi2WO6 NAs, NiFe-LDH/Bi2WO6 NAs, Ti3C2/Bi2WO6 NAs, and NiFe-LDH/Ti3C2/Bi2WO6 NAs.
Nanomaterials 14 00477 g006
Figure 7. Schematic representation illustrating the photoelectrocatalytic degradation mechanism of organic compounds via the NiFe-LDH/Ti3C2/Bi2WO6 nanocomposite photoelectrode. This diagram delineates the process of photogenerated electron–hole pair separation and subsequent transfer, highlighting the synergistic effects of the composite materials in enhancing the degradation efficiency of organic pollutants (The red and green arrows indicate the direction of movement of electrons and holes, respectively).
Figure 7. Schematic representation illustrating the photoelectrocatalytic degradation mechanism of organic compounds via the NiFe-LDH/Ti3C2/Bi2WO6 nanocomposite photoelectrode. This diagram delineates the process of photogenerated electron–hole pair separation and subsequent transfer, highlighting the synergistic effects of the composite materials in enhancing the degradation efficiency of organic pollutants (The red and green arrows indicate the direction of movement of electrons and holes, respectively).
Nanomaterials 14 00477 g007
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, Y.; Li, R.; Zhang, J.; Liu, L.; Huang, W.; Wang, Y. LDH/MXene Synergistic Carrier Separation Effects to Improve the Photoelectric Catalytic Activities of Bi2WO6 Nanosheet Arrays. Nanomaterials 2024, 14, 477. https://doi.org/10.3390/nano14050477

AMA Style

Wang Y, Li R, Zhang J, Liu L, Huang W, Wang Y. LDH/MXene Synergistic Carrier Separation Effects to Improve the Photoelectric Catalytic Activities of Bi2WO6 Nanosheet Arrays. Nanomaterials. 2024; 14(5):477. https://doi.org/10.3390/nano14050477

Chicago/Turabian Style

Wang, Yuting, Runhua Li, Jiaying Zhang, Liming Liu, Weiwei Huang, and Yajun Wang. 2024. "LDH/MXene Synergistic Carrier Separation Effects to Improve the Photoelectric Catalytic Activities of Bi2WO6 Nanosheet Arrays" Nanomaterials 14, no. 5: 477. https://doi.org/10.3390/nano14050477

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