Next Article in Journal
Microenvironmental Behaviour of Nanotheranostic Systems for Controlled Oxidative Stress and Cancer Treatment
Next Article in Special Issue
The Effect of La3+ on the Methylene Blue Dye Removal Capacity of the La/ZnTiO3 Photocatalyst, a DFT Study
Previous Article in Journal
Controllable Valley Polarization and Strain Modulation in 2D 2H–VS2/CuInP2Se6 Heterostructures
Previous Article in Special Issue
N, P Self-Doped Porous Carbon Material Derived from Lotus Pollen for Highly Efficient Ethanol–Water Mixtures Photocatalytic Hydrogen Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ag3PO4-Deposited TiO2@Ti3C2 Petals for Highly Efficient Photodecomposition of Various Organic Dyes under Solar Light

Department of Mechanical Engineering, Gachon University, Seongnam 13120, Korea
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(14), 2464; https://doi.org/10.3390/nano12142464
Submission received: 1 June 2022 / Revised: 9 July 2022 / Accepted: 9 July 2022 / Published: 18 July 2022

Abstract

:
Two-dimensional Ti3C2 MXenes can be used to fabricate hierarchical TiO2 nanostructures that are potential photocatalysts. In this study, the photodecomposition of organic dyes under solar light was investigated using flower-like TiO2@Ti3C2, deposited using narrow bandgap Ag3PO4. The surface morphology, crystalline structure, surface states, and optical bandgap properties were determined using scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), nitrogen adsorption analysis, and UV-Vis diffuse reflectance spectroscopy (UV-DRS). Overall, Ag3PO4-deposited TiO2@Ti3C2, referred to as Ag3PO4/TiO2@Ti3C2, demonstrated the best photocatalytic performance among the as-prepared samples, including TiO2@Ti3C2, pristine Ag3PO4, and Ag3PO4/TiO2 P25. Organic dyes, such as rhodamine B (RhB), methylene blue (MB), crystal violet (CV), and methylene orange (MO), were efficiently degraded by Ag3PO4/TiO2@Ti3C2. The significant enhancement of photocatalysis by solar light irradiation was attributed to the efficient deposition of Ag3PO4 nanoparticles on flower-like TiO2@Ti3C2 with the efficient separation of photogenerated e-/h+ pairs, high surface area, and extended visible-light absorption. Additionally, the small size of Ag3PO4 deposition (ca. 4–10 nm diameter) reduces the distance between the core and the surface of the composite, which inhibits the recombination of photogenerated charge carriers. Free radical trapping tests were performed, and a photocatalytic mechanism was proposed to explain the synergistic photocatalysis of Ag3PO4/TiO2@Ti3C2 under solar light.

1. Introduction

Globally, as industrialization and urbanization accelerate, the demand for food and consumer goods (clothing, electronics, furniture, vehicles, etc.) increases. Unsustainable resource use and inefficient waste management, however, contribute to environmental problems, which have become a source of concern [1,2,3]. Numerous harmful pollutants endanger the environment and human health, such as heavy metal ions, toxic gases, and organic dye [4,5]. The majority of organic dyes found in wastewater originate from various industrial waste streams, including paint, dyeing clothing, bleaching paper, synthesizing rubber, and processing plastic, thereby, resulting in water pollution.
Organic dyes at high concentrations (5–1500 mg/L) cause significant environmental and organismal toxicity, including carcinogenic, mutagenic, and teratogenic effects [6,7,8]. Owing to their low removal capacity, conventional biological processes used to remove dyes, such as flocculation, filtration, precipitation, and coagulation, have gradually become obsolete [9,10]. In comparison, heterogeneous photocatalysts used in advanced oxidation processes are considered promising alternatives for the removal of a wide variety of organic pollutants with a highly efficient degradation rate [11].
Titania (TiO2) is the most widely used and recognized photocatalytic material in environmental remediation owing to its superior photocatalytic performance, non-toxicity, and stability [12,13,14]. However, pure TiO2 has a low quantum efficiency and a large bandgap, limiting its application (i.e., electrons can be triggered only under UV light). In addition, the rapid recombination of photogenerated electron–hole pairs reduces its degradation efficiency [15]. Recently, several strategies have been proposed to solve these problems by fabricating TiO2 derived from Ti3C2 [16].
Despite the absence of noble metal deposition (Au, Ag, Pt, Ru2O, etc.), the flower-shaped TiO2@Ti3C2 composite exhibits significantly enhanced photocatalytic activity [4,17]. Recent work has been published on the synthesis of TiO2@Ti3C2 nanoflowers from the MAX phase (Ti3AlC2) [4,16,17]. They demonstrated that the presence of Ti3C2 in the TiO2 structure eliminates electron–hole pair recombination, maximizes charging transfer, and expands the light absorption area to use visible light. Under solar light irradiation, the degradation efficiencies of RhB were 97% within 40 min and 95% within 60 min. Therefore, flower-shaped TiO2@Ti3C2 is a highly promising photocatalyst [4,16,17].
Silver-based semiconductor photocatalysts have been widely used due to their high efficiency as visible-light-driven photocatalysts [18]. Among silver-based photocatalysts, trisilver phosphate (Ag3PO4) with a narrow bandgap (2.36 eV) has attracted the photocatalytic field’s attention due to its high oxidation capacity and ability to remove pollutants under visible light illumination [18,19,20,21]. Nevertheless, Ag3PO4 continues to fall short of meeting the demand for large-scale industrial applications, primarily due to its low reuse rates, as it suffers greatly from severe photocorrosion [22,23].
This is due to the intrinsic fast charge recombination, self-corrosion, and photocorrosion of silver or silver oxides in the absence of a sacrificial reagent. Therefore, the key strategy is to prevent its photocorrosion in practical applications. In this regard, heterostructure photocatalysts composed of Ag3PO4 and TiO2 semiconductors have attracted attention over the years due to their increased photocatalytic efficiency, increased stability, and lower noble metal consumption [20]. To our knowledge, no research on the combination of Ag3PO4 and Ti3C2-derived TiO2 has been conducted.
Herein, an Ag3PO4-deposited TiO2@Ti3C2 composite, Ag3PO4/TiO2@Ti3C2, was developed as a highly efficient visible photocatalyst. Flower-like TiO2 was initially synthesized from Ti3C2 MXenes through three consecutive steps of hydrothermal oxidation, ion exchange, and heating processes [4,16,17]. Thereafter, using an in-situ precipitation method, Ag3PO4 nanoparticles (NPs) were deposited on flower-like TiO2@Ti3C2. It is conceivable that the presence of metal-like Ti3C2 in the TiO2@Ti3C2 composite acts as an “electron sink”, facilitating the highly efficient photodegradation of organic dyes.
In addition, the flower-like morphology has a high surface area and an efficient deposition of Ag3PO4 on TiO2@Ti3C2 petals, which effectively separates electron–hole pairs and improves the solar light harvesting capability. Ag3PO4-deposited TiO2@Ti3C2 exhibited significant photocatalytic performance in the photodecomposition of a variety of organic dyes (including RhB, MB, CV, and MO).

2. Materials and Methods

2.1. Chemical and Materials

Titanium carbide powder (Ti3C2 MXenes) was provided by Invisible Co. Ltd., Seoul, Korea. Sodium hydroxide (NaOH) was purchased from Daejung Chemicals & Metals Co., Ltd., Siheung, Korea. Silver nitrate (AgNO3, 99.9%) was obtained from Duksan Pure Chemicals Co., Ltd., Ansan, Korea. Hydrogen peroxide (H2O2, 30%), hydrochloric acid (HCl), ethyl alcohol (C2H5OH), sodium phosphate (Na3PO4), methylene blue (MB), rhodamine B (RhB), methylene orange (MO), and crystal violet (CV) were purchased from Sigma-Aldrich, Munich, Germany. All chemicals were used directly, without any further treatment.

2.2. Preparation of TiO2@Ti3C2 Heterostructure

The procedure for fabricating flower-like TiO2@Ti3C2 was previously described by Vu Thi Quyen et al. [17]. First, 100 mg of Ti3C2 MXene was added and vigorously stirred for 15 min with a solution of 2 M NaOH (80 mL) and 30 wt% H2O2 (8 mL). Second, the hydrothermal reaction was performed by transferring the mixture into two 50 mL autoclave systems at 140 °C for 15 h. The dispersion was then allowed to cool naturally to room temperature, and the samples were washed with deionized (DI) water and ethanol several times before being dried in an oven at 60 °C for 12 h. The dried sample was soaked in a 0.05 M HCl solution (500 mL) for 12 h to ensure ion exchange. After heating the sample in a furnace at 500 °C for 5 h, the TiO2@Ti3C2 composite was obtained. The formation of TiO2 is presented by the following equations:
2H2O2 → 2H2O + O2
Ti3C2 + 5O2 → 3TiO2 + 2CO2
3TiO2 + 2NaOH → Na2Ti3O7 + H2O
Na2Ti3O7 + HCl → H2Ti3O7 + 2NaCl
H2Ti3O7 → 3TiO2 + H2O

2.3. Preparation of Ag3PO4 Particles

To prepare Ag3PO4 particles, a simple method was followed. To generate Ag3PO4, 0.1 M Na3PO4 (8 mL) was added dropwise to 0.1 M AgNO3 solution (24 mL) and stirred for 5 h in the dark. To remove redundant ions, the precipitate was centrifuged and washed with deionized water. The purified sample was dried overnight at 60 °C to obtain the Ag3PO4 powder.

2.4. Preparation of Ag3PO4/TiO2@Ti3C2

Ag3PO4/TiO2@Ti3C2 composites were prepared by adding various amounts of AgNO3 (x) and Na3PO4 (y) solutions with a ratio of x:y = 3:1. First, 0.05 g TiO2@Ti3C2 was added to 0.1 M AgNO3 solutions. The mixed solutions were vigorously stirred and sonicated for 10 min. Thereafter, 0.1 M Na3PO4 was gradually dropped into the solution and stirred for 5 h to form Ag3PO4. To avoid photocorrosion, the reaction was conducted in the dark. The resulting powder products were washed, centrifuged with DI water, and dried at 60 °C to obtain Ag3PO4/TiO2@Ti3C2. The powder obtained from mixing quantities x = 1, 2, 4, 6, and 8 mL AgNO3 with y = 0.33, 0.67, 1.33, 2, and 2.67 mL Na3PO4 corresponds to samples marked A1 to A5, respectively). The samples were kept in the dark throughout the preparation process to avoid photocorrosion. In comparison with sample A4, Ag3PO4/TiO2 P25 was also prepared under the same conditions.
The fabrication procedure for the composites is depicted in Scheme 1. First, Ti3C2 MXene was subjected to hydrothermal oxidation, ion exchange, and heat treatment steps, transforming its accordion-like structure into a flower shape. Thereafter, Ag3PO4 NPs containing varying amounts of Ag3PO4 were deposited by in-situ precipitation on the flower-like structure to form the Ag3PO4/TiO2@Ti3C2 composites. Photodegradation tests were conducted under solar-driven light.

2.5. Photocatalysis of Dye Degradation

A Xe lamp (1000 W) was used as an artificial solar light source with a light intensity of 1000 mW/cm2. Under simulated solar light, the photocatalytic activities of the as-prepared samples were evaluated for the photodecomposition of RhB dye. Typically, 10 mg of each sample was added to 20 mL of an aqueous solution containing RhB (9 mg/L). Prior to light irradiation, the adsorption experiments of Ag3PO4, TiO2, and Ag3PO4/TiO2@Ti3C2 were performed by stirring the solutions in the dark for 15 min to improve dispersion and adsorption–desorption equilibrium. Thereafter, the solution was exposed to solar light. Following that, an aliquot of each sample (1 mL) was removed and centrifuged at the indicated intervals to obtain the supernatant for UV-vis spectrophotometer evaluation.
The degradation efficiency of the dye materials can be described by the following equation: degradation percentage (%) = Co − Ct/Co × 100. To determine the rate constant of RhB degradation, the degradation kinetics were assumed to follow the pseudo-first order model [21,24]:
l n C C 0 = k t
where Co and Ct denote the initial concentration of RhB and a specific concentration of RhB after exposure to light for t minutes, respectively. Here, “k” is the pseudo-first-order rate constant calculated from the linear slope of ln(C0/C) versus t (time). 10 mg/L concentration of MB, MO, and CV were used for photodegradation with the same procedure. The absorbance changes for each dye were determined using a UV-Vis spectrometer at different wavelengths, RhB (554 nm), MB (664–665 nm), MO (464 nm), and CV (590 nm).

2.6. Characterization

The X-ray diffraction (XRD) patterns of the samples were determined using a Rigaku Smartlab X-ray diffractometer with Cu-Kα radiation (λ = 1.544 Å) (Rigaku Corporation, Tokyo, Japan). The morphologies and microstructures of the samples were investigated using a Hitachi S-4700 field emission scanning electron microscope (FE-SEM) (Hitachi Ltd., Tokyo, Japan) and FEI Tecnai transmission electron microscopy (TEM) (FEI, Hillsboro, OR, United States). XPS measurements were conducted using an X-ray photoelectron spectrometer (XPS, Multilab 2000, Thermo Scientific, Waltham, MA, United States). The surface area, pore size, and pore volume were measured using an N2 adsorption–desorption apparatus (ASAP 2020, Micromeritics Instrument Corp, Norcross, GA, USA). Optical properties of the samples were determined using a Jasco V770 UV-Vis diffuse reflectance spectrophotometer (Jasco Inc., Easton, MD, USA)

3. Results and Discussion

3.1. XRD Analysis

The diffraction patterns of the TiO2@Ti3C2, Ag3PO4, and A4 samples are shown in Figure 1a. The XRD data of Ag3PO4 revealed sharp and narrow dominant peaks at 33.3° (210) and 36.6° (211), which corresponded to the body-centered crystal structure and high crystallinity (JCPDS no. 06-0505) [25]. Furthermore, TiO2@Ti3C2 exhibited a distinct peak at 2θ = 25.3°, indicating the high crystallinity of the anatase phase with other weaker peaks at 37.8°, 53.9°, 55.34°, and 62.6° (JCPDS card No. 21-1272) [21]. A small peak at 48.5° confirmed the presence of Ti3C2 in the sample [4,16,17].
In the case of the composites, Figure 1b shows the diffraction patterns of Ag3PO4/TiO2@Ti3C2 (samples A3, A4, and A5), which clearly display both Ag3PO4 crystal and TiO2 anatase phase peaks, confirming that the composites were composed of Ag3PO4 and TiO2. The diffraction peaks of TiO2 in the heterostructure composite are weaker than those of Ag3PO4 owing to the lower crystallinity of TiO2. However, samples A1 and A2 containing a lower concentration of Ag3PO4 mostly exhibit peaks of TiO2 anatase at = 25.3° (JCPDS card No. 21-1272), instead of Ag3PO4 peaks due to self-corrosion [22].

3.2. Morpholoy Analysis by SEM, TEM

The accordion-like shape of Ti3C2 is shown in the SEM image in Figure 2a. After oxidization, ion exchange, and heat treatment, accordion-shape Ti3C2 transforms into TiO2@Ti3C2 with a flower shape (Figure 2b). Figure 2c depicts pure Ag3PO4 NPs with an irregular shape and a diameter of ca. 300–500 nm [19]. The SEM image in Figure 2d shows aggregated Ag3PO4-deposited TiO2@Ti3C2 that retains its flower-like shape. To overcome certain limitations of SEM analysis, which focuses on the surface morphology of samples, TEM and HRTEM images were used to clarify the transmission morphology and crystalline structure of Ag3PO4/TiO2@Ti3C2 (sample A4). According to the EDX spectrum (Figure S1), the main elements present in the sample are carbon, oxygen, titanium, silver, and phosphorous. C, O, and Ti were obtained from TiO2@Ti3C2, with a portion of the O derived from Ag3PO4. It can be demonstrated that Ag3PO4 is coated on the surface of TiO2@Ti3C2.
Figure 3a shows a typical TEM image of the flower-shaped TiO2@Ti3C2 with some petal fragments and large agglomerated Ag3PO4 on it. However, the higher magnification images in Figure 3b,c demonstrate that the dominant Ag3PO4 NPs were formed and deposited on each TiO2@Ti3C2 nanorod. Positively charged silver ions were attracted to the surface of negatively charged TiO2, and the reaction between Ag+ and PO43− occurred immediately upon the addition of Na3PO4 solution, resulting in the deposition of Ag3PO4 NPs on TiO2@Ti3C2 rods [26]. The TEM images show that precipitating Ag3PO4 using TiO2@Ti3C2 nanorods as a template can reduce particle aggregation owing to the high specific surface area of the nanorods, which affects Ag3PO4 nucleation and reduces the diameter to around 4–10 nm.
Additionally, the Ag3PO4 nanoparticles were relatively stable and did not detach during sonication as part of the TEM preparation process, whereas the agglomerated Ag3PO4 fragments moved and changed position during real-time TEM measurements. Figure 3c,d show the high-resolution TEM (HR-TEM) images of the Ag3PO4/TiO2@Ti3C2 composite. TiO2@Ti3C2 (101) has lattice fringes with an interplanar spacing of 0.35 nm, while cubic Ag3PO4 (210) has an interplanar spacing of 0.26 nm (Figure 3d). A SAED image further demonstrates the clear formation and polycrystallinity of these materials, which is shown in Figure S2. These TEM images can assist in resolving the initial ambiguity regarding the morphology of Ag3PO4/TiO2@Ti3C2 shown in the SEM images.
Figure 4 depicts the scanning TEM (STEM) image of Ag3PO4/TiO2@Ti3C2 with elemental mapping. The high-angle annular dark-field (HAADF) STEM image of Ag3PO4/TiO2@Ti3C2 shows areas with different contrasts, with the brighter regions representing Ag3PO4 particles and the darker regions representing TiO2@Ti3C2 rods. The elemental mapping analysis displays C, O, P, K, and Ag elements from an arbitrary area. In detail, the Ti signal is primarily associated with the backbone of TiO2@Ti3C2, whereas the P and Ag signals are associated with the large agglomerated Ag3PO4 fragments on top of the flower and Ag3PO4 NPs decorated alongside the petals. Additionally, the O signals are distributed uniformly throughout the composite. The elemental mapping results provide more comprehensive and clear observations of the elemental distribution across the whole composite, corresponding to the SEM and HR-TEM above.

3.3. XPS Results

To characterize the chemical compositions and elemental states of the as-prepared samples, XPS measurements were conducted. The survey spectra of Ti3C2 MXene, TiO2@Ti3C2 flowers, and Ag3PO4/TiO2@Ti3C2 illustrate the characteristic elemental peaks at approximately 532.35, 455.16, and 284.96 eV allocated to O 1s, Ti 2p, and C 1s, respectively (Figure S3). The weak peak of F 1s at 685.1 eV in Ti3C2, indicates that some fluoride was left over from the synthesis process [17].
The TiO2@Ti3C2 heterostructure exhibits a much stronger O 1s peak intensity when compared to Ti3C2, indicating the existence of an appreciable amount of oxide in the heterostructure due to the oxidation process and the formation of TiO2 [4]. In the spectra results of composite Ag3PO4/TiO2@Ti3C2, the peaks of P 2p and Ag 3d represent Ag3PO4 in the composite. Generally, Ag 3d peaks have a higher intensity than Ti 2p, suggesting that many silver elements or their compounds are present on the surface of the solid sample, decorating the TiO2@Ti3C2 flower shape. This result is consistent with the SEM and TEM images discussed previously.
Figure 5 demonstrates the Ti 2p, C 1s, and O 1s high-resolution spectra of Ti3C2, TiO2@Ti3C2, and Ag3PO4/TiO2@Ti3C2 (sample A4). The Ti-C peak at 454.26 eV in Ti 2p spectra indicates the presence of Ti3C2; however, there is no Ti-C peak in TiO2@Ti3C2 [27,28]. Additionally, the intensification of the Ti-O peak (Ti4+) at 458.83 eV indicates the formation of TiO2 in the flower-shaped TiO2@Ti3C2. The Ti-C bond is absent in TiO2@Ti3C2 in the high-resolution C 1s; however, C–C/C–H, C–O, and O–C = O bonds exist at approximately 284.88, 286.40, and 288.72 eV, respectively [28,29,30]. Owing to the use of TiO2@Ti3C2 as a template for precipitate Ag3PO4, the Ti 2p and C 1s spectra of Ag3PO4/TiO2@Ti3C2 (sample A4) are similar to those of TiO2@Ti3C2.
Nevertheless, the differences in O 1s XPS spectra among the three samples were unavoidable owing to the changing environment in which oxygen was present. Ti3C2 exhibited a distinct peak attributed to C-Ti-OH linkage at 530.72 and 532.59 eV, whereas the others showed the strongest Ti-O peak at 529.60 eV. Furthermore, Ti3C2 MXene also exhibited F 1s spectra (Figure S4a), which could be ascribed to residual fluorine after the etching steps. The Ag 3d and P 1s spectra of Ag3PO4/TiO2@Ti3C2 (sample A4) are presented in Figure S4b,c.
The binding energies of 366.83 and 372.91 eV assigned to Ag 3d5/2 and Ag 3d3/2, respectively, indicate that Ag+ is dominated in the composites. The deconvoluted Ag 3d peaks are slightly shifted toward lower binding energy, and the FHMW is also wider. This is attributed to self-corrosion after exposure to the environment for a period of time [22]. The peak of the P 2p spectra located at 131.82 eV corresponds to P5+ in the PO43+ [31].

3.4. Surface Area and Optical Analysis

The N2 adsorption–desorption isotherms and the corresponding BET surface areas, pore volumes, and pore sizes of TiO2@Ti3C2, Ag3PO4/P25, and Ag3PO4/TiO2@Ti3C2 are shown in Figure 6a and Table 1. Based on the isotherm curves, all samples belong to type IV, and the pore size indicates that they are mesopores [4]. As can be seen, the TiO2@Ti3C2 flowers have the highest BET surface area (53 m2g−1). After the deposition of Ag3PO4, the surface area and pore volume of the Ag3PO4/TiO2@Ti3C2 sample (40 m2g−1) decreased slightly but remained higher than that of Ag3PO4/TiO2 P25 (27 m2g−1). This high surface area enables photocatalytic reactions to occur.
The light-harvesting capability is important for evaluating photocatalytic activity as demonstrated by UV-Vis absorption spectra in Figure 6b. Owing to its metallic properties, Ti3C2 MXene exhibits a nearly horizontal spectrum over the entire wavelength range. The sequential transformations of MXene (metallic material) to the TiO2@Ti3C2 heterostructure (semiconductor material) throughout the processes resulted in the gradual development of the TiO2@Ti3C2 absorption peak [4,16,17]. Its light absorption is improved compared with that of the commercial TiO2 P25.
It can be seen that the decoration of Ag3PO4 on the flower-shaped TiO2@Ti3C2 extends the optical absorption by the combination of Ag3PO4 and TiO2@Ti3C2. Due to the high surface area of the 3D nanoflower structure, the harvested and scattered light is omnidirectional. In this case, the forward and backward light results in constructive interference, which can extend the photon lifetime and improve the absorbance [16]. These findings further confirm that the Ag3PO4-deposited TiO2@Ti3C2 composite can enhance the light-harvesting ability, making it a promising photocatalyst under visible illumination.

3.5. Photocatalytic Performance

3.5.1. The Effects of Ag3PO4 Content on the Photodegradation of Rhodamine B

Figure 7a illustrates the photocatalytic degradation of RhB (9 mgL−1) using Ag3PO4, TiO2 and Ag3PO4/TiO2@Ti3C2 composites (samples A1, A2, A3, A4, and A5) at a concentration of 0.5 gL−1 under solar light irradiation. Overall, the composite exhibited effective photodegradation reaction under solar light irradiation in a short period of time, with sample A4 displaying significant enhanced photocatalytic performance by demonstrating 97% RhB degradation within 12 min.
As samples A1 and A2 contain less Ag3PO4 and may be susceptible to self-corrosion and photocorrosion, as discussed in the XRD diffractogram, these samples took 20 min to partially degrade 28.9% and 59.9% RhB, respectively. Although sample A5 contains more Ag3PO4 than sample A4, there was no significant difference in the photocatalytic activities between samples A5 and A4. When comparing degradation performances within 20 min, pure Ag3PO4 had a degradation efficiency of 95% and Ag3PO4/TiO2 P25 86%. The lower photocatalytic performance of Ag3PO4/TiO2 P25 is most likely due to self-corrosion in the composite after prolonged storage [22].
Figure 7b shows the calculated rate constant (k) values of 0.017, 0.046, 0.161, 0.289, and 0.278; 0.223; 0.095 min−1; and 0.0093 min−1 for samples A1–A5, Ag3PO4, Ag3PO4/TiO2 P25, and flower-like TiO2@Ti3C2, respectively. These results indicate that, among the as-prepared samples, sample A4 exhibited the greatest photocatalysis performance. These results indicate that combining TiO2 derived from MXene with Ag3PO4 can significantly improve its photocatalytic performance when exposed to solar illumination.

3.5.2. Photodegradation of Other Organic Dyes

The three dyes, MB, CV, and MO (10 mg L−1, 0.5 gL−1) were also photodegraded under solar light irradiation. The adsorption and photocatalytic performance of sample A4 over time for the three organic dyes are shown in Figure 8b and Figure S5. The results indicate that the composite is effective in adsorbing cationic dyes (MB and CV) but has no effect on anionic dye concentration (MO) after 15 min in the dark. It can be attributed to the composite’s surface, which is anionic and negatively charged in the aqueous solution, which enables it to readily attract and adsorb cationic dyes in comparison to its anionic dye counterpart [32]. These results aid in the cationic dyes’ molecules more easily reaching the surface of materials, leading to higher degradation performance in MB and CV compared to MO. After 6 min, the photocatalyst degraded 94.4% of MB, nearly 99.2% CV in 14 min, and 92.4% MO after 40 min. The pseudo-first-order rate constants for MB, CV, and MO were calculated to be 0.489, 0.279, and 0.073 min−1, respectively.
Compared with TiO2@Ti3C2, pristine Ag3PO4, and Ag3PO4/TiO2 P25, the Ag3PO4/TiO2@Ti3C2 composite exhibited superior photocatalytic activity. Recent publications on photocatalytic materials indicate the possibility of a highly promising photocatalyst for flower shaped TiO2@Ti3C2 with a high surface area. Combining TiO2@Ti3C2 with Ag3PO4 results in synergetic photodegradation of various dyes. The flower-like structure of TiO2 derived from Ti3C2 can improve its light-harvesting ability, and its high specific surface area allows the solvent to approach the reactive sites more easily.
Additionally, Ag3PO4 in the composites can shorten the diffusion paths for photoexcited carriers, thereby, inhibiting the recombination of electron–hole pairs [4,25]. The photocatalytic activity of the optimized sample was compared to that of other reported similar photocatalysts. As demonstrated in Table 2, the Ag3PO4/TiO2@Ti3C2 composite (sample A4) can degrade a variety of pollutants, including both cationic and anionic dyes, in a relatively short period of time, as compared to the state-of-the-art works listed in Table 2.

3.6. Scavenger Trapping and Recycling Tests of RhB Degradation

To ascertain the main active species in the photocatalytic reactions, scavenger trapping tests were conducted on RhB photodegradation over the Ag3PO4/TiO2@Ti3C2 composite (sample A4). The three types of scavengers were tert-butanol (a quencher of •OH), Na2-EDTA (a quencher of h+), and p-benzoquinone (p-BQ, a quencher of •O2−). The dye solution containing a specified amount of catalyst was added to 2 mL of trapping agent (0.01 M), stirred for 15 min in the dark, and exposed to solar light for 20 min. As shown in Figure 9a, the addition of tert-butanol had no effect on the photocatalytic performance, thereby, indicating that free •OH radicals were not the dominant oxidizing species. The presence of Na2-EDTA and p-BQ, however, decreased the degradation efficiency to 4.76% and 21.7%, respectively. From these results, it can be concluded that h+ and •O2 radicals play a critical role in the photodecomposition of RhB.
Along with the photocatalytic efficiency, stability evaluations of photocatalysts are necessary for practical use. Figure 9b shows the RhB degradation efficiency of sample A4 over three consecutive recycling runs. The degradation efficiency was significantly reduced after repeated exposure to solar light, indicating that the photocatalyst was not stable and rapidly deteriorated due to the photocorrosion of Ag3PO4 NPs [22]. In the first run, a degradation efficiency of 97% was obtained in 12 min, whereas the degradation efficiencies of RhB were approximately 74% and 39% after the second and third runs, respectively. After a few recycling tests under solar light, the photocorrosion of the catalyst resulted in a low photodecomposition rate by forming an uncontrolled amount of Ag0, which can agglomerate and hinder the photocatalytic ability [31].

3.7. Proposed Photocatalytic Mechanism

Based on the obtained scavenger trapping and recycling results, the photocatalytic mechanism of Ag3PO4/TiO2@Ti3C2 is schematically shown in Figure 10. Under solar illumination, electrons from the valence band (VB) of Ag3PO4 could be excited to its conduction band (CB), resulting in the formation of holes in the valence band (VB). Since the EVB of TiO2 is more negative than that of Ag3PO4, holes from Ag3PO4 also transferred to TiO2 and directly degraded RhB, which was absorbed on the surface of TiO2 during the oxidization process [25,31]. According to literature, the conduction potential of Ag3PO4 is higher than the activation energy of single-electron oxygen, preventing photogenerated electrons from being captured by dissolved oxygen.
Instead, a small amount of Ag is formed as a result of photocorrosion [31]. Electrons and holes in Ag could also be excited by light energy, with electrons migrating to the CB of TiO2, while the remaining holes could be recombined with photoexcited electrons in Ag3PO4 CB, thereby, partially preventing further corrosion [31]. Furthermore, some electrons were transferred from TiO2 and Ag3PO4 to Ti3C2 MXene, resulting in band bending and the formation of a Schottky junction as well as a uniform Fermi level [4,17,39]. The photo-reduction reaction occurring on the surface of Ti3C2 and the TiO2 conduction band generated •O2 radicals, which are required for effective RhB decomposition. The production of •O2 and h+ could easily degrade a variety of dyes, including RhB, MB, CV, and MO, under solar light. The proposed mechanism is consistent with scavenger trapping experiments and is formulated as follows:
  • Photo-oxidation reaction:
Ag3PO4 + hv → Ag3PO4 (h+ + e)
Ag3PO4 (h+) + TiO2 → Ag3PO4 + TiO2 (h+)
TiO2 (h+) + RhB → By-products + CO2 + H2O
2.
Photo-reduction reaction:
  • Photoexcited electrons and separation of e/h+:
Ag+ + Ag3PO4 (e) → Ag + Ag3PO4
Ag + hv → Ag (h+ + e)
  • Transfer routes and formation of O2●−:
Ag3PO4 (e) + Ag (h+) → Ag3PO4 + Ag
Ag (e) + TiO2 → Ag + TiO2 (e)
TiO2 (e) + Ti3C2 → TiO2 + Ti3C2 (e)
Ag3PO4 (e) + Ti3C2 → Ag3PO4 + Ti3C2 (e)
Ti3C2 (e) + O2 → Ti3C2 + O2●−
TiO2 (e) + O2 → TiO2 + O2●−
  • Degradation of pollutants:
O2●− + RhB → By-products + CO2 + H2O
Based on this proposed mechanism, the stability of the composite did not significantly improve. This might be due to a longer electron transfer distance between TiO2 and Ti3C2 than between Ag3PO4 and Ti3C2. Second, an unknown amount of Ti3C2 on the surface and/or insufficient interfacial contact between Ag3PO4 and Ti3C2 might lower the composite’s stability. Both reasons could not prevent the rapid formation of Ag, reducing the stability and photocatalytic performance of the composite.

4. Conclusions

This study successfully synthesized Ag3PO4/TiO2@Ti3C2 composites by precipitating Ag3PO4 NPs on the surface of TiO2@Ti3C2 flowers. Some characterization methods were conducted to investigate the surface morphology, structural composition, surface area, and optical properties of Ag3PO4/TiO2@Ti3C2. The characteristics of Ag3PO4-deposited TiO2@Ti3C2 endow superior photocatalytic activities in the comparison with TiO2@Ti3C2, pristine Ag3PO4, and Ag3PO4/TiO2 P25, particularly sample A4. This composite exhibited excellent photocatalytic performance for various organic dyes (degraded 97% RhB, 94% MB, 99% CV, and 92% MO) within a short period of time when exposed to solar light irradiation.
The high photocatalytic activity of the composite was due to the high surface area of TiO2@Ti3C2 with small Ag3PO4 particles (4–10 nm) that can easily reach dye molecules. The e/h+ transfer throughout the composite system also contributes to increased charging transfer, expanded light absorption wavelength, and decreased electron–hole pair recombination, which gives synergistic effects for photocatalytic activity. Based on the scavenger trapping and recycling test results, the mechanism was proposed.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano12142464/s1, Figure S1: EDS patterns of Ag3PO4/TiO2@Ti3C2 (sample A4); the inset image shows an SEM image of the sample, Figure S2: SAED image of Ag3PO4/TiO2@Ti3C2 (sample A4) taken of an arbitrary area, Figure S3: XPS survey spectra of Ti3C2, TiO2@Ti3C2, and Ag3PO4/TiO2@Ti3C2, Figure S4: (a) F 1s high-resolution XPS spectra of Ti3C2, and (b–c) Ag 3d and P 1s high-resolution XPS spectra of Ag3PO4/TiO2@Ti3C2 (sample A4), Figure S5: Photodegradation absorption spectra of (a) Methylene Blue, (b) Crystal Violet, and (c) Methylene Orange under solar light.

Author Contributions

Conceptualization, N.T.A.N.; methodology, N.T.A.N.; software, N.T.A.N.; validation, N.T.A.N. and H.K.; formal analysis, N.T.A.N.; investigation, N.T.A.N.; resources, N.T.A.N. and H.K.; data curation, N.T.A.N. and H.K.; writing—original draft preparation, N.T.A.N.; writing—review and editing, N.T.A.N. and H.K.; visualization, N.T.A.N.; supervision, H.K.; project administration, H.K.; funding acquisition, H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded in part by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade, Industry & Energy (MOTIE) of the Republic of Korea (No. 20203010040010) and in part by the Korea Evaluation Institute of Industrial Technology (KEIT) and the Ministry of Trade, Industry & Energy (MOTIE) of the Republic of Korea (No. 200153646).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

The authors deeply appreciate Professor Sang-Wha Lee at Gachon University for his valuable advice on this study.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Li, P.; Karunanidhi, D.; Subramani, T.; Srinivasamoorthy, K. Sources and Consequences of Groundwater Contamination. Arch. Environ. Contam. Toxicol. 2021, 80, 1–10. [Google Scholar] [CrossRef] [PubMed]
  2. Schweitzer, L.; Noblet, J. Water Contamination and Pollution. In Green Chemistry—An Inclusive Approach; Elsevier: Amsterdam, The Netherlands, 2018; pp. 261–290. [Google Scholar] [CrossRef]
  3. My Tran, N.; Thanh Hoai Ta, Q.; Sreedhar, A.; Noh, J.S. Ti3C2Tx MXene playing as a strong methylene blue adsorbent in wastewater. Appl. Surf. Sci. 2021, 537, 148006. [Google Scholar] [CrossRef]
  4. My Tran, N.; Thanh Hoai Ta, Q.; Noh, J.S. Unusual synthesis of safflower-shaped TiO2/Ti3C2 heterostructures initiated from two-dimensional Ti3C2 MXene. Appl. Surf. Sci. 2021, 538, 148023. [Google Scholar] [CrossRef]
  5. Muneer, I.; Farrukh, M.A.; Ali, D.; Bashir, F. Heterogeneous photocatalytic degradation of organic dyes by highly efficient GdCoSnO3, Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2021, 265, 115028. [Google Scholar] [CrossRef]
  6. Ton, N.N.T.; Dao, A.T.N.; Kato, K.; Ikenaga, T.; Trinh, D.X.; Taniike, T. One-pot synthesis of TiO2/graphene nanocomposites for excellent visible light photocatalysis based on chemical exfoliation method. Carbon 2018, 133, 109–117. [Google Scholar] [CrossRef]
  7. Kadam, R.L.; Kim, Y.; Gailkwad, S.; Chang, M.; Tarte, N.H.; Han, S. Catalytic decolorization of rhodamine B, Congo red, and crystal violet dyes, with a novel niobium oxide anchored molybdenum (Nb–O–Mo). Catalysts 2020, 10, 491. [Google Scholar] [CrossRef]
  8. Chiu, Y.H.; Chang, T.F.M.; Chen, C.Y.; Sone, M.; Hsu, Y.J. Mechanistic insights into photodegradation of organic dyes using heterostructure photocatalysts. Catalysts 2019, 9, 430. [Google Scholar] [CrossRef] [Green Version]
  9. Ghaffar, A.; Zhang, L.; Zhu, X.; Chen, B. Porous PVdF/GO Nanofibrous Membranes for Selective Separation and Recycling of Charged Organic Dyes from Water. Environ. Sci. Technol. 2018, 52, 4265–4274. [Google Scholar] [CrossRef]
  10. Pal, B.; Kaur, R.; Grover, I.S. Superior adsorption and photodegradation of eriochrome black-T dye by Fe3+ and Pt4+ impregnated TiO2 nanostructures of different shapes. J. Ind. Eng. Chem. 2016, 33, 178–184. [Google Scholar] [CrossRef]
  11. Sacco, O.; Vaiano, V.; Rizzo, L.; Sannino, D. Photocatalytic activity of a visible light active structured photocatalyst developed for municipal wastewater treatment. J. Clean. Prod. 2018, 175, 38–49. [Google Scholar] [CrossRef]
  12. Etacheri, V.; Di Valentin, C.; Schneider, J.; Bahnemann, D.; Pillai, S.C. Visible-light activation of TiO2 photocatalysts: Advances in theory and experiments. J. Photochem. Photobiol. C Photochem. Rev. 2015, 25, 1–29. [Google Scholar] [CrossRef] [Green Version]
  13. Daghrir, R.; Drogui, P.; Robert, D. Modified TiO2 for environmental photocatalytic applications: A review. Ind. Eng. Chem. Res. 2013, 52, 3581–3599. [Google Scholar] [CrossRef]
  14. Wang, Y.; He, Y.; Lai, Q.; Fan, M. Review of the progress in preparing nano TiO2: An important environmental engineering material. J. Environ. Sci. 2014, 26, 2139–2177. [Google Scholar] [CrossRef] [PubMed]
  15. Ansari, S.A.; Cho, M.H. Highly Visible Light Responsive, Narrow Band gap TiO2 Nanoparticles Modified by Elemental Red Phosphorus for Photocatalysis and Photoelectrochemical Applications. Sci. Rep. 2016, 6, 25405. [Google Scholar] [CrossRef]
  16. Li, Y.; Deng, X.; Tian, J.; Liang, Z.; Cui, H. Ti3C2 MXene-derived Ti3C2/TiO2 nanoflowers for noble-metal-free photocatalytic overall water splitting. Appl. Mater. Today 2018, 13, 217–227. [Google Scholar] [CrossRef]
  17. Quyen, V.T.; Ha, L.T.T.; Thanh, D.M.; van Le, Q.; Viet, N.M.; Nham, N.T.; Thang, P.Q. Advanced synthesis of MXene-derived nanoflower-shaped TiO2@Ti3C2 heterojunction to enhance photocatalytic degradation of Rhodamine B. Environ. Technol. Innov. 2021, 21, 101286. [Google Scholar] [CrossRef]
  18. Frontistis, Z.; Antonopoulou, M.; Petala, A.; Venieri, D.; Konstantinou, I.; Kondarides, D.I.; Mantzavinos, D. Photodegradation of ethyl paraben using simulated solar radiation and Ag3PO4 photocatalyst. J. Hazard. Mater. 2017, 323, 478–488. [Google Scholar] [CrossRef]
  19. Zhao, F.M.; Pan, L.; Wang, S.; Deng, Q.; Zou, J.J.; Wang, L.; Zhang, X. Ag3PO4/TiO2 composite for efficient photodegradation of organic pollutants under visible light. Appl. Surf. Sci. 2014, 317, 833–838. [Google Scholar] [CrossRef]
  20. Xie, J.; Yang, Y.; He, H.; Cheng, D.; Mao, M.; Jiang, Q.; Song, L.; Xiong, J. Facile synthesis of hierarchical Ag3PO4/TiO2 nanofiber heterostructures with highly enhanced visible light photocatalytic properties. Appl. Surf. Sci. 2015, 355, 921–929. [Google Scholar] [CrossRef]
  21. Khalid, N.R.; Mazia, U.; Tahir, M.B.; Niaz, N.A.; Javid, M.A. Photocatalytic degradation of RhB from an aqueous solution using Ag3PO4/N-TiO2 heterostructure. J. Mol. Liq. 2020, 313, 113522. [Google Scholar] [CrossRef]
  22. Ma, X.; Li, H.; Wang, Y.; Li, H.; Liu, B.; Yin, S.; Sato, T. Substantial change in phenomenon of “self-corrosion” on Ag3PO4/TiO2 compound photocatalyst. Appl. Catal. B Environ. 2014, 158–159, 314–320. [Google Scholar] [CrossRef]
  23. Chen, X.; Dai, Y.; Wang, X. Methods and mechanism for improvement of photocatalytic activity and stability of Ag3PO4: A review. J. Alloys Compd. 2015, 649, 910–932. [Google Scholar] [CrossRef]
  24. Nekouei, F.; Nekouei, S.; Pouzesh, M.; Liu, Y. Porous-CdS/Cu2O/graphitic-C3N4 dual p-n junctions as highly efficient photo/catalysts for degrading ciprofloxacin and generating hydrogen using solar energy. Chem. Eng. J. 2020, 385, 123710. [Google Scholar] [CrossRef]
  25. Nyankson, E.; Efavi, J.K.; Agyei-Tuffour, B.; Manu, G. Synthesis of TiO2-Ag3PO4photocatalyst material with high adsorption capacity and photocatalytic activity: Application in the removal of dyes and pesticides. RSC Adv. 2021, 11, 17032–17045. [Google Scholar] [CrossRef]
  26. Zanella, R.; Giorgio, S.; Henry, C.R.; Louis, C. Alternative methods for the preparation of gold nanoparticles supported on TiO2. J. Phys. Chem. B 2002, 106, 7634–7642. [Google Scholar] [CrossRef]
  27. Kong, F.; He, X.; Liu, Q.; Qi, X.; Zheng, Y.; Wang, R.; Bai, Y. Improving the electrochemical properties of MXene Ti3C2 multilayer for Li-ion batteries by vacuum calcination. Electrochim. Acta 2018, 265, 140–150. [Google Scholar] [CrossRef]
  28. Cao, M.; Wang, F.; Wang, L.; Wu, W.; Lv, W.; Zhu, J. Room Temperature Oxidation of Ti3C2 MXene for Supercapacitor Electrodes. J. Electrochem. Soc. 2017, 164, A3933–A3942. [Google Scholar] [CrossRef]
  29. Shen, C.; Wang, L.; Zhou, A.; Wang, B.; Wang, X.; Lian, W.; Hu, Q.; Qin, G.; Liu, X. Synthesis and electrochemical properties of two-dimensional RGO/Ti3C2Tx nanocomposites. Nanomaterials 2018, 8, 80. [Google Scholar] [CrossRef] [Green Version]
  30. Xia, Q.X.; Fu, J.; Yun, J.M.; Mane, R.S.; Kim, K.H. High volumetric energy density annealed-MXene-nickel oxide/MXene asymmetric supercapacitor. RSC Adv. 2017, 7, 11000–11011. [Google Scholar] [CrossRef] [Green Version]
  31. Liu, W.; Liu, D.; Wang, K.; Yang, X.; Hu, S.; Hu, L. Fabrication of Z-scheme Ag3PO4/TiO2 Heterostructures for Enhancing Visible Photocatalytic Activity. Nanoscale Res. Lett. 2019, 14, 203. [Google Scholar] [CrossRef] [Green Version]
  32. Azeez, F.; Al-Hetlani, E.; Arafa, M.; Abdelmonem, Y.; Nazeer, A.A.; Amin, M.O.; Madkour, M. The effect of surface charge on photocatalytic degradation of methylene blue dye using chargeable titania nanoparticles. Sci. Rep. 2018, 8, 7104. [Google Scholar] [CrossRef] [PubMed]
  33. Li, Y.; Yu, L.; Li, N.; Yan, W.; Li, X. Heterostructures of Ag3PO4/TiO2 mesoporous spheres with highly efficient visible light photocatalytic activity. J. Colloid Interface Sci. 2015, 450, 246–253. [Google Scholar] [CrossRef] [PubMed]
  34. Tang, C.; Liu, E.; Fan, J.; Hu, X.; Kang, L.; Wan, J. Heterostructured Ag3PO4/TiO2 nano-sheet film with high efficiency for photodegradation of methylene blue. Ceram. Int. 2014, 40, 15447–15453. [Google Scholar] [CrossRef]
  35. Pant, B.; Prasad Ojha, G.; Acharya, J.; Park, M. Ag3PO4-TiO2-Carbon nanofiber Composite: An efficient Visible-light photocatalyst obtained from electrospinning and hydrothermal methods. Sep. Purif. Technol. 2021, 276, 119400. [Google Scholar] [CrossRef]
  36. Saud, P.S.; Pant, B.; Twari, A.P.; Ghouri, Z.K.; Park, M.; Kim, H.Y. Effective photocatalytic efficacy of hydrothermally synthesized silver phosphate decorated titanium dioxide nanocomposite fibers. J. Colloid Interface Sci. 2016, 465, 225–232. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Duoerkun, G.; Shi, Z.; Cao, W.; Liu, T.; Liu, J.; Zhang, L.; Li, M.; Chen, Z. Construction of TiO2/Ag3PO4 nanojunctions on carbon fiber cloth for photocatalytically removing various organic pollutants in static or flowing wastewater. J. Colloid Interface Sci. 2020, 571, 213–221. [Google Scholar] [CrossRef]
  38. Sheu, F.J.; Cho, C.P.; Liao, Y.T.; Yu, C.T. Ag3PO4-TiO2-graphene oxide ternary composites with efficient photodegradation, hydrogen evolution, and antibacterial properties. Catalysts 2018, 8, 57. [Google Scholar] [CrossRef] [Green Version]
  39. Cai, T.; Wang, L.; Liu, Y.; Zhang, S.; Dong, W.; Chen, H.; Yi, X.; Yuan, J.; Xia, X.; Liu, C.; et al. Ag3PO4/Ti3C2 MXene interface materials as a Schottky catalyst with enhanced photocatalytic activities and anti-photocorrosion performance. Appl. Catal. B Environ. 2018, 239, 545–554. [Google Scholar] [CrossRef]
Scheme 1. Schematic indicating the fabrication processes, beginning with multilayer Ti3C2 MXene and ending with the formation of Ag3PO4-deposited flower-shaped TiO2@Ti3C2.
Scheme 1. Schematic indicating the fabrication processes, beginning with multilayer Ti3C2 MXene and ending with the formation of Ag3PO4-deposited flower-shaped TiO2@Ti3C2.
Nanomaterials 12 02464 sch001
Figure 1. XRD patterns of as-prepared samples: (a) TiO2@Ti3C2, Ag3PO4, and optimized A4 sample. (b) Additional XRD patterns of A1, A2, A3, and A5 samples, including A4.
Figure 1. XRD patterns of as-prepared samples: (a) TiO2@Ti3C2, Ag3PO4, and optimized A4 sample. (b) Additional XRD patterns of A1, A2, A3, and A5 samples, including A4.
Nanomaterials 12 02464 g001
Figure 2. SEM images of (a) Ti3C2, (b) TiO2@Ti3C2, (c) Ag3PO4, and (d) Ag3PO4/TiO2@Ti3C2 (A4) samples.
Figure 2. SEM images of (a) Ti3C2, (b) TiO2@Ti3C2, (c) Ag3PO4, and (d) Ag3PO4/TiO2@Ti3C2 (A4) samples.
Nanomaterials 12 02464 g002
Figure 3. (ac) TEM images of Ag3PO4/TiO2@Ti3C2 (sample A4) viewed under different magnifications. (d) HR-TEM lattice image of Ag3PO4/TiO2@Ti3C2.
Figure 3. (ac) TEM images of Ag3PO4/TiO2@Ti3C2 (sample A4) viewed under different magnifications. (d) HR-TEM lattice image of Ag3PO4/TiO2@Ti3C2.
Nanomaterials 12 02464 g003
Figure 4. STEM images of Ag3PO4/TiO2@Ti3C2 (sample A4) with element mapping (C, O, P, Ti, and Ag).
Figure 4. STEM images of Ag3PO4/TiO2@Ti3C2 (sample A4) with element mapping (C, O, P, Ti, and Ag).
Nanomaterials 12 02464 g004
Figure 5. Ti 2p, C 1s, and O 1s high-resolution XPS spectra of Ti3C2 (ac), TiO2@Ti3C2 (df), and Ag3PO4/TiO2@Ti3C2 (A4) (gi).
Figure 5. Ti 2p, C 1s, and O 1s high-resolution XPS spectra of Ti3C2 (ac), TiO2@Ti3C2 (df), and Ag3PO4/TiO2@Ti3C2 (A4) (gi).
Nanomaterials 12 02464 g005
Figure 6. (a) BET surface area plot of TiO2@Ti3C2, Ag3PO4/P25, and Ag3PO4/TiO2@Ti3C2; (b) UV-Vis DRS spectra of the as-prepared samples, including pristine TiO2 P25, TiO2@Ti3C2, pure Ag3PO4, and Ag3PO4/TiO2@Ti3C2 (sample A4).
Figure 6. (a) BET surface area plot of TiO2@Ti3C2, Ag3PO4/P25, and Ag3PO4/TiO2@Ti3C2; (b) UV-Vis DRS spectra of the as-prepared samples, including pristine TiO2 P25, TiO2@Ti3C2, pure Ag3PO4, and Ag3PO4/TiO2@Ti3C2 (sample A4).
Nanomaterials 12 02464 g006
Figure 7. (a) Photocatalytic performance and (b) the rate constant of Ag3PO4/TiO2@Ti3C2 composite (samples A1, A2, A3, A4, and A5), pure Ag3PO4, TiO2@Ti3C2 (B1), and Ag3PO4/TiO2 P25 (B2).
Figure 7. (a) Photocatalytic performance and (b) the rate constant of Ag3PO4/TiO2@Ti3C2 composite (samples A1, A2, A3, A4, and A5), pure Ag3PO4, TiO2@Ti3C2 (B1), and Ag3PO4/TiO2 P25 (B2).
Nanomaterials 12 02464 g007
Figure 8. (a) Photocatalytic results of Ag3PO4/TiO2@Ti3C2 (sample A4) and (b) the corresponding rate constants for organic dyes.
Figure 8. (a) Photocatalytic results of Ag3PO4/TiO2@Ti3C2 (sample A4) and (b) the corresponding rate constants for organic dyes.
Nanomaterials 12 02464 g008
Figure 9. (a) Effects of scavengers on the RhB photodecomposition. (b) The stability of Ag3PO4/TiO2@Ti3C2 (A4) after three consecutive recycling runs.
Figure 9. (a) Effects of scavengers on the RhB photodecomposition. (b) The stability of Ag3PO4/TiO2@Ti3C2 (A4) after three consecutive recycling runs.
Nanomaterials 12 02464 g009
Figure 10. A schematic illustration of the proposed photocatalytic mechanism of Ag3PO4/TiO2@Ti3C2 (A4) exposed to solar light.
Figure 10. A schematic illustration of the proposed photocatalytic mechanism of Ag3PO4/TiO2@Ti3C2 (A4) exposed to solar light.
Nanomaterials 12 02464 g010
Table 1. Specific surface area, pore volume, and pore size distribution of TiO2@Ti3C2, Ag3PO4/TiO2 P25, and Ag3PO4/TiO2@Ti3C2 (sample A4).
Table 1. Specific surface area, pore volume, and pore size distribution of TiO2@Ti3C2, Ag3PO4/TiO2 P25, and Ag3PO4/TiO2@Ti3C2 (sample A4).
SampleBET Surface Area
(m2g−1)
Pore Volume (cm3g−1)Pore Size (nm)
TiO2@Ti3C2530.30921.91
Ag3PO4/TiO2 P25270.20231.04
Ag3PO4/TiO2@Ti3C2 (A4)400.15620.12
Table 2. A comparison of the photocatalytic efficiency of TiO2@Ti3C2 and Ag3PO4-based photocatalysts developed for dye degradation.
Table 2. A comparison of the photocatalytic efficiency of TiO2@Ti3C2 and Ag3PO4-based photocatalysts developed for dye degradation.
PollutantPollutant conc.PhotocatalystDecoration Ag3PO4 Method Ag3PO4 DiameterDosage (g L−1)Irradiation Time (min)Efficiency (%)Light Source[Ref.]
RhB
MO
MB
CV
9 ppm
10 ppm
10 ppm
10 ppm
Ag3PO4/TiO2@Ti3C2In situ precipitation4–10 nm0.512
40
6
14
97
92.4
94
99.2
Solar lightCurrent study
RhB20 µmSafflower-shaped TiO2/Ti3C2--0.86095Visible light[4]
RhB10 ppmTiO2@Ti3C2 nanoflower--0.434097Visible light[17]
RhB,
MO,
phenol
0.02 mM
0.06 mM
0.2 mM
Ag3PO4/TiO2 In situ precipitation300–500 nm0.515
80
90
~100
90
~70
Visible light [19]
RhB10 ppmAg3PO4/TiO2 NFsIn situ precipitation300–360 nm0.410 (Vis)~100Visible light[20]
RhB,
MB,
pesticides
6 ppm
8 ppm
-
Ag3PO4/TiO2One pot100–400 nm0.5~ 6
4
98
99.1
Tungsten halogen light[25]
MB20 ppmAg3PO4/m-TiO2 In situ precipitation~4 nm0.528100Visible light [33]
MB10 mMAg3PO4/TiO2 nanosheet filmImpregnating-deposition process100–300 nm-7098.1Visible light [34]
MB10 ppmAg3PO4-TiO2-CNFsHydrothermal20–32 nm0.510100Visible light[35]
MB10 ppmAg3PO4/TiO2 NFsHydrothermal-0.415100Sunlight[36]
RhB10 ppmCFC/TiO2/Ag3PO4Hydrothermal20–100 nm-10098.4Visible light[37]
MO20 ppmAg3PO4-TiO2-GOIon exchange250–600 nm112080Solar light[38]
RhB10 ppmAg3PO4/TiO2Hydrothermal and ultrasonication0.1–1.5 µm125100Solar light[31]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nguyen, N.T.A.; Kim, H. Ag3PO4-Deposited TiO2@Ti3C2 Petals for Highly Efficient Photodecomposition of Various Organic Dyes under Solar Light. Nanomaterials 2022, 12, 2464. https://doi.org/10.3390/nano12142464

AMA Style

Nguyen NTA, Kim H. Ag3PO4-Deposited TiO2@Ti3C2 Petals for Highly Efficient Photodecomposition of Various Organic Dyes under Solar Light. Nanomaterials. 2022; 12(14):2464. https://doi.org/10.3390/nano12142464

Chicago/Turabian Style

Nguyen, Ngoc Tuyet Anh, and Hansang Kim. 2022. "Ag3PO4-Deposited TiO2@Ti3C2 Petals for Highly Efficient Photodecomposition of Various Organic Dyes under Solar Light" Nanomaterials 12, no. 14: 2464. https://doi.org/10.3390/nano12142464

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