Next Article in Journal
The Impact of Technology on Adolescent Sexual and Reproductive Needs
Next Article in Special Issue
Heavy Metal Pollution in Xinfengjiang River Sediment and the Response of Fish Species Abundance to Heavy Metal Concentrations
Previous Article in Journal
Changes in the Work Schedule of Nurses Related to the COVID-19 Pandemic and Its Relationship with Sleep and Turnover Intention
Previous Article in Special Issue
Comparative Study of Algal Responses and Adaptation Capability to Ultraviolet Radiation with Different Nutrient Regimes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of High-Porosity B-TiO2/C3N4 Composite Materials: Adsorption–Degradation Capacity and Photo-Regeneration Properties

1
College of Environment, Hohai University, Nanjing 210098, China
2
College of Mechanics and Materials, Hohai University, Nanjing 211100, China
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2022, 19(14), 8683; https://doi.org/10.3390/ijerph19148683
Submission received: 24 June 2022 / Revised: 13 July 2022 / Accepted: 15 July 2022 / Published: 17 July 2022

Abstract

:
Adsorption can quickly remove pollutants in water, while photocatalysis can effectively decompose organic matter. B-TiO2/g-C3N4 ternary composite photocatalytic materials were prepared by molten method, and their adsorption–degradation capability under visible light conditions was discussed. The morphology of the B-TiO2/g-C3N4 materials was inspected by SEM, TEM, BET, and EDS, and the results showed that close interfacial connections between TiO2 and g-C3N4, which are favorable for charge transfer between these two semiconductors, formed heterojunctions with suitable band structure which was contributed by the molten B2O3. Meanwhile, the molten B2O3 effectively increased the specific surface area of TiO2/C3N4 materials, thereby increasing the active sites and reducing the recombination of photogenerated electron–hole pairs and improving the photocatalytic degradation abilities of TiO2 and g-C3N4. Elsewhere, the crystal structure analysis (XRD, XPS, FTIR) results indicated that the polar -B=O bond formed a new structure with TiO2 and g-C3N4, which is not only beneficial for inhibiting the recombination of electron holes but also improving the photocatalytic activity. By removal experiment, the adsorption and degradation performances of B-TiO2/g-C3N4 composite material were found to be 8.5 times and 3.4 times higher than that of g-C3N4. Above all, this study prepared a material for removing water pollutants with high efficiency and provides theoretical support and experimental basis for the research on the synergistic removal of pollutants by adsorption and photocatalysis.

1. Introduction

Photocatalysis is a technology that can use semiconductor materials to remove pollutants from the environment under different lighting conditions [1,2,3]. Most materials require ultraviolet conditions to produce redox effects. The visible part of sunlight irradiate that reaches the water surface only occupies 45%, while the ultraviolet part occupies less than 4% [4]. At the same time, water absorbs and reflects sunlight in different wavelengths, which seriously limits the removal efficiency of photocatalytic materials in a water environment [5,6,7].
g-C3N4 is a good semiconductor material with application potential that is metal free and has visible light responsiveness [8,9,10]. However, researchers found that g-C3N4 still has many disadvantages such as low utilization of visible light, poor photoelectric conversion efficiency, and low specific surface area [11]. Since Serpone et al. [12] first reported that a solid–solid heterojunction interface with good contact can be constructed from different coupled semiconductor materials to promote electron transfer between particles, more and more researchers have paid attention to different semiconductor materials. Among the many photocatalytic semiconductor materials, TiO2 is widely used in sewage treatment [13], photocatalytic synthesis [14], and self-cleaning [15]. Due to its easy availability, low cost, stable chemical properties, corrosion resistance, non-toxicity, and strong oxidizing properties [16], its air purification [17] and antibacterial properties [18] have been extensively studied. Therefore, using the advantages of TiO2 material properties to composite it with g-C3N4 to become a more competitive material has attracted the attention of more and more researchers [19,20,21].
In the treatment of water environment pollution, the adsorption performance and visible light response ability of photocatalytic materials are two important factors that determine whether the photocatalytic technology can be effectively promoted [22,23]. The high-efficiency adsorption and enrichment ability of the material reduce the concentration of pollutants in water and provide a high-concentration contact environment that is conducive to the photocatalytic reaction for the material [24,25]. The melting characteristics of B2O3-modified g-C3N4 are effective for improving its specific surface area and adsorption performance, but its response to visible light is limited, which affects the visible light photocatalytic activity [26].
Effectively improving the adsorption performance and visible light catalytic activity of materials at the same time is a key issue for researchers. Functional coupling through different materials is an idea to solve this problem. Here, TiO2 and g-C3N4 were formed into a composite heterostructure to improve the photocatalytic oxidation ability of the material. Then, the melting characteristics of B2O3 during the heating process were used as a “reaction environment regulator”, and TiO2/C3N4 was co-calcined to prepare a composite material. The results of the experiment showed that B element doped the composite-modified materials, and, finally, the adsorption behavior and visible light catalytic degradation ability of B-TiO2/C3N4 were analyzed by using MB and RhB organic pollutants. This method effectively increased the specific surface area of the material, increased the conjugated system, improved the adsorption capacity, and effectively improved the visible light catalytic effect of the material. This provides a technical idea for the promotion and application of a low-cost preparation of an efficient adsorption–degradation photocatalytic composite material.

2. Materials and Methods

2.1. Sample Preparation

(1)
TiO2: 17 mL of butyl titanate (Aladdin, 99% pure) was placed into 55 mL of ethanol solution (Aladdin, 99.7% pure), then 4.5 mL of glacial acetic acid (Aladdin, 99% pure) was added. After mixing using a magnetic stirrer, the solution was recorded as A. We measured 27.6 mL of ethanol, added 0.9 mL of distilled water, and adjusted the pH to 4 with nitric acid (Sinopharm, 65−68% pure), and this solution was recorded as B after thorough mixing. The solutions A and B were stirred for 30 min respectively, the B solution was added dropwise to the A solution at a rate of 10 drops per minute, and the final obtained mixed solution was TiO2 sol. Then, the TiO2 sol was stirred at room temperature and placed in an oven to dry after gelation. Then, the dried xerogel was ground into powder and placed in a crucible with a lid to heat at a rate of 5 °C/min until 550 °C and was maintained at this temperature for 2 h. After the temperature in the muffle furnace (XS2−10, Li Chen, China) dropped to room temperature, the calcined material was ground into powder, and the obtained material was denoted as TiO2.
(2)
g-C3N4: 10 g melamine (Aladdin, 99% pure) was placed into a 50 mL crucible and put into muffle furnace, and the heating rate was set to 5 °C/min until it reached the reaction temperature of 550 °C. Then, the reaction temperature was maintained for 2 h. After the material cooled to room temperature, we ground the calcined solid, and the obtained yellow powder was g-C3N4.
(3)
TiO2/C3N4 composite material: 1 g of melamine and 10 mL of TiO2 sol were weighed in a beaker, stirred, and mixed thoroughly and then the mixture was placed into an oven for drying after gelation. The dried composite precursor was placed in a crucible with a lid and calcined at 550 °C for 2 h. The obtained material was ground into powder after cooling to room temperature to obtain TiO2/C3N4 composite material.
(4)
B-C3N4: The B2O3 (Aladdin, 99.9% pure) and g-C3N4 were mixed and ground evenly. Then, the composite powder was placed into a ceramic crucible. After that, the samples were heated to 550 °C at a heating rate of 5 °C/min and maintained for 2 h. When the furnace was cooled to room temperature, the powders were washed with water and ethanol several times. For convenience of description, the composite material was abbreviated as B-C3N4.
(5)
B-TiO2/C3N4: The TiO2/C3N4 material and B2O3 were mixed and stirred in distilled water at a mass ratio of 1:1. After grinding evenly, we put it into an atmosphere furnace at 550 °C for 2 h. Then, the material was soaked and washed three times with ethanol and distilled water, respectively. The sample was denoted as B-TiO2/C3N4. For comparison, another ternary composite material, B2-C3N4/TiO2, was prepared: 1 g of B-C3N4 and 10 mL of TiO2 sol were weighed with the same method.

2.2. Characterization

The structural analysis of the samples was carried out by X-ray diffraction (XRD, Rigaku Ultima Ⅲ, Tokyo, Japan) and was recorded in the 2θ range of 5–80° with a scan rate of 0.02°/0.4 s using a Bruker AXSD8 system (Bruker, Billerica, MA, USA) equipped with a Cu Kα radiation source (λ = 0.15406 Å, in which the X-ray tube was operated at 40 kV and 40 mA). UV–Vis diffuse reflectance spectra (DRS) were obtained on a UV–visible (UV–Vis) spectrophotometer (PerkinElmer, Waltham, MA, USA) with BaSO4 as the reference. The sample morphology of the different composite materials was examined using a scanning electron microscope (SEM, Hitachi, Tokyo, Japan) and a transmission electron microscope (TEM, JEM-2100, JEOL, Tokyo, Japan). X-ray photoelectron spectroscopy (XPS) measurements were performed on a Thermo Scientific ESCALAB 250 instrument (Thermo Fisher Scientific, Waltham, MA, USA) with an Al Kα source. Low-temperature N2 adsorption/desorption measurements (Brunauer–Emmett–Teller (BET) method) were carried out using a Micromeritics ASAP 2020 system (Micromeritics, Norcross, GA, USA) at −196 °C following degassing of all samples at 120 °C for 2 h.

2.3. Adsorption and Photocatalytic Degradation of Organic Pollution

In order to evaluate whether composite materials have a better effect on the removal of pollutants, TiO2, g-C3N4, B-TiO2, B-C3N4, TiO2/C3N4, B-TiO2/C3N4, and B2-C3N4/TiO2 were weighed in the removal experimental. A 30 mg amount of each of material was added to the MB solution with a volume of 100 mL and a concentration of 20 mg/L. Under the condition of magnetic stirring, the photocatalytic efficiency was measured after 30 min of dark reaction adsorption experiment. A 1.5 mL amount of solution was taken out for measurement each time, and the removal rate of pollutants was determined by the following formula:
R t = C 0 C t C 0 × 100 %
where Rt is the removal rate at time t after commencing the adsorption and photocatalytic degradation process, and C0 and Ct are initial concentration and concentration at time t, respectively.
Q e = C 0 C e m V
where Qe is the adsorbed quantity of samples at the equilibrium moment of adsorption and desorption. C0, Ce, V, and m are the initial concentration, concentration at time t, initial volume of MB, and quantity of adsorbent, respectively.

2.4. Kinetics and Adsorption Isotherm Model

The adsorption behaviors were fitted to the materials by pseudo-first-order kinetic and pseudo-second-order kinetic models, respectively. Two kinetic linear models are shown below [27]:
Pseudo-first-order kinetic model:
l n ( Q e Q t ) = l n Q e k 1 t
Pseudo-second-order kinetic model:
t Q t = 1 k 2 Q e 2 + t Q e
where Qt and Qe in the formula are the adsorption amount of the sample at time t and equilibrium, respectively. k1 and k2 are pseudo-first-order kinetic constants and pseudo-second-order kinetic constants, respectively.
Langmuir and Freundlich are the two most common adsorption isotherm models. The Langmuir model assumes that the surface of the adsorbate is uniform and the adsorption capacity is the same everywhere, and it only occurs on the outer surface of the adsorbent; this is monolayer adsorption. The Freundlich adsorption equation can be applied to both monolayer adsorption and adsorption behavior on uneven surfaces. Freundlich, as an empirical adsorption isotherm for non-uniform surfaces, is more applicable to low-concentration adsorption and can interpret experimental results over a wider concentration range. Therefore, in this experiment, the experimental results are fitted by these two adsorption isotherm models, and the formulas are as follows [28]:
Langmuir:
C e Q e = 1 Q m K L + C e Q m
Freundlich:
l n Q e = 1 n l n C e + l n K F
where Qe and Qm represent the adsorption capacity and the maximum adsorption capacity, respectively; Ce is the concentration of pollutants in the solution at adsorption equilibrium; KL is the Langmuir adsorption equilibrium constant; and KF is the Freundlich constants with the affinity coefficient.

3. Result and Discussion

3.1. Performance Testing

In order to study the physical and chemical properties of the materials, the materials were screened by the adsorption catalytic performance of removing pollutant MB (Figure S1). It can be clearly seen from the figure that the adsorption–catalytic efficiency of several different materials for removing MB under the dark reaction and visible light conditions was B-TiO2/C3N4 > B-C3N4 > B2-C3N4/TiO2 > TiO2/C3N4 > B-TiO2 > g-C3N4 > TiO2. Among them, the adsorption and removal rate of MB on B-TiO2/C3N4 in the dark reaction process reached 73.8%, while that of B2-C3N4/TiO2 was only 17.8%. Then, under the visible light condition for 2 h, the removal rate of MB by B-TiO2/C3N4 reached 97.3%, while that of B2-C3N4/TiO2 only reached 66.5%. Therefore, according to the preliminary experimental results, the B-TiO2/C3N4 material was selected as the composite photocatalyst with the highest removal efficiency for follow-up research.
The mixing ratio and calcination temperature of the composites are both important factors affecting the performance of the composite materials. Therefore, in order to further analyze the performance of the composite material, this experiment first selected B2O3 and TiO2/C3N4 with different mass ratios to be mixed and calcined at 550 °C and then selected the material with the best pollutant removal effect for subsequent experiments.
According to the amount of doping of B2O3, the composite materials were marked as B-TiO2/C3N4-x (x = 1, 2, 3, 4, 5). The removal efficiencies are shown in Figure S2. The removal rates of B-TiO2/C3N4-1, B-TiO2/C3N4-2, and B-TiO2/C3N4-3 were 69.3%, 83.2%, and 73.4%, respectively, while the removal rates of B-TiO2/C3N4-4 and B-TiO2/C3N4-5 were only 50.9% and 23.9%.
Therefore, according to the above experimental results, B-TiO2/C3N4-2 had the best effect and was selected as the follow-up research material. A 2 g amount of B2O3 and 1 g of TiO2/C3N4 material were mixed uniformly, and the mixture temperature was heated at a rate of 5 °C/min for 2 h at different temperatures. The set target temperatures were 350 °C, 450 °C, 550 °C, 650 °C, and 750 °C. For convenience of description, B-TiO2/C3N4 prepared at different calcination temperatures was named after its temperature (350 °C, 450 °C, 550 °C, 650 °C, and 750 °C).

3.2. Characterization and Analysis of B-TiO2/C3N4 Composites

The morphology of B-TiO2/C3N4 composites prepared under different calcination conditions at different temperatures are shown in Figure 1. It can be seen from Figure 1a that at 350 °C the surface of the structure was relatively smooth, which may have been due to molten B2O3 wrapping the base material. Figure 1b shows the surface layer of the calcined material at 450 °C was still partially covered, but a large number of loose structures also appeared. This may have been due to the increase in the calcination temperature, and the disorder degree increased due to the B2O3 entering the molten state, which contains polar -B=O groups. As seen in Figure 1c, many loose pores appeared on the surface of the composite under the 550 °C calcination condition. As the temperature continued to increase, the main structure of the g-C3N4 material began to decompose when the temperature was above 600 °C, and the pulverization of the g-C3N4 structure can be clearly seen in Figure 1d. At 750 °C, g-C3N4 was decomposed into nitrogen and nitrile-based fragments, and TiO2 was gradually transformed from anatase to rutile; so, the layered structure of g-C3N4 cannot be observed in Figure 1e, while some agglomerates of particles can be seen.
The specific surface area and porosity of materials are crucial factors affecting the adsorption and photocatalytic reactions. As shown in Figure S3, the representative N2 adsorption–desorption isotherms of composite materials were of type IV [29]. According to the classification of IUPAC, the adsorption and desorption curve hysteresis loop types of the prepared B-TiO2/C3N4 materials belonged to the “H3” hysteresis loop (0.5 < P/P0 < 0.98:550 °C, 650 °C, 750 °C; 0.75 < P/P0 < 0.97:350 °C, 450 °C). The low-temperature calcination shown for loops located at high P/P0 was caused by the aggregation of the 2D lamellar structure of g-C3N4. By contrast, the high temperature and B2O3 as a medium effectively resulted in more holes on the surface of g-C3N4. In Table 1, details regarding the specific surface areas, average pore diameters, and pore volumes of various materials are summarized. Among the temperature series, 550 °C showed the highest specific surface areas, average pore diameters, and pore volumes. This is because, at the calcination temperature of 550 °C, the molten B2O3 could fully contact the TiO2/C3N4 composite as a reaction environment modifier. The lamellar structure of g-C3N4 was exfoliated in the molten environment while the pore structure was formed during the cooling process, so the specific surface area was significantly increased.
The surface morphologies and microstructure of B-TiO2/C3N4 (550 °C) composite material are shown in Figure 2. From Figure 2a,b, it can be clearly observed that there were a lot of loose macropores and particle agglomeration on the surface of the g-C3N4-based material. In order to better analyze the morphological characteristics of the material, the high-resolution transmission electron microscope (HR-TEM) pattern was used for analysis. In Figure 2c, it can be seen that there were a lot of loose pores on the surface of the g-C3N4. Otherwise, a large number of granular material lattice fringes appeared on the g-C3N4 lamellar structure. The lattice spacing embedded particles were measured to be ~0.35 nm (Figure 2d), which is similar to the TiO2 (101) crystal plane structure [30]. Through the analysis of specific surface area and the characterization of SEM and TEM, it was found that TiO2 grew uniformly on the surface of the g-C3N4 layered structure, and the modification of the TiO2/C3N4 composite structure by B2O3 produced a large number of loose pores on the g-C3N4 layered structure. Such a modification method not only improved the specific surface area of the material, but also preserved the composite structure of TiO2/C3N4, which is more conducive to the charge transfer of photogenerated carriers between composite semiconductor materials and improves the photocatalytic activity of the material.
The XRD pattern of materials was recorded and is shown in Figure S4. It can be seen that B2O3 showed two obvious, characteristic peaks at 14.6°and 27.9° [31]. The composites prepared at 350 °C and 450 °C had only two obvious characteristic peaks of B2O3 but no characteristic peaks of TiO2 and g-C3N4 [32]. This may have been because the surface of the TiO2/C3N4 material was coated by B2O3 in the molten state at 350 °C and 450 °C, which is basically consistent with the SEM characterization shown in Figure 1. At the same time, due to the large amount of B2O3 added, there was no peak position of g-C3N4 and TiO2 but only the characteristic peak of B2O3. At 550 °C, the two characteristic peaks of B2O3 disappeared in the XRD pattern, and the anatase (25.3°) and rutile (27.4°) diffraction peaks of TiO2 appeared. It was caused by the entry of polar -B=O into the TiO2/C3N4. When the temperature rose to 650 °C, several characteristic peaks (27.4°, 41.3°, etc.) of the rutile phase of TiO2 became more and more obvious, which indicated that the crystal structure of TiO2 gradually changed from anatase to rutile with the increase in temperature. It can be seen from XRD analysis that when the temperature gradually increased, the degree of disorder gradually increased, and a polar -B=O bond was formed to form a new structure with TiO2 or g-C3N4 [33].
The Fourier-transform infrared spectroscopy of composite materials is shown in Figure S5. In this figure, it can be seen that the peak at 550 °C was the smallest, which was the Ti–O bond expansion joint and Ti stretching vibration of the -O–Ti bond. Compared with the B-TiO2/C3N4 composites at different calcination temperatures (350 °C, 450 °C, 550 °C, 650 °C, and 750 °C), the obvious peaks appeared at 810 cm−1 and 1200–1600 cm−1. The characteristic peak at 810 cm−1 was the stretching vibration of the triazine ring structure in g-C3N4, while the broad peak spectrum at 1200–1600 cm−1 was caused by the stretching vibration of C=N. Otherwise, the obvious differences of all composite materials were mainly due to the incomplete transformation of B2O3 from solid to molten state. At low temperatures, the peaks were not obvious, which was caused by molten B2O3 coating on the surface of composite material, resulting in limited exposure of the surface structure of the B-TiO2/C3N4 composite. In addition, the absorption peaks of B-TiO2/C3N4 composites at 3000–3300 cm−1 were mainly the stretching vibration of the amino group (-NH2). Meanwhile, the two peaks at 2260 cm−1 and 2350 cm−1 were the C=O stretching vibration of CO2 in the atmosphere [34].
Figure S6 shows the XPS spectra of several elements of B-TiO2/C3N4: B1s, N1s, C1s, Ti2s and O1s. B1s shows three binding energies at 190.4 eV, 192.7 eV, and 193.5 eV, of which 192.7 eV belonged to the B–O bond, 190.4 eV was the binding energy of the B–N bond, and 193.5 eV may have been O–B formed by B entering the TiO2 crystal structure—Ti or TiB2 bond [35]. This indicates that B element entered into the structures of TiO2 and g-C3N4. C1s showed three main binding energies, of which 284.8 eV was the typical sp2 hybridization of the graphite phase (C=C bond peak position), and 288.7 eV corresponded to the C–N–C bond in the structure. In addition, 286.5 eV corresponded to the exocyclic C–O of carbon–nitrogen polymer materials [36]. Figure S6c shows the N1s binding energy peaks at 397.3 eV, 398.2 eV, and 401.9 eV. Among them, 398.2 eV was attributed to the sp2 hybridization (C–N=C) of the N triazine ring structure, while the peak at 401.9 eV corresponded to the bridged N atom (N–(C)3) [37], and the peak at 397.3 eV corresponded to N–B [38]. In addition, the characteristic peaks in the N spectrum were slightly shifted, possibly due to the increased delocalization between the large π bonds in the carbon nitride structure due to the doping of B element [29]. The two peaks of 458.7 eV and 464.8 eV in the binding energy of Ti 2s were the electron binding energies of Ti4+ 2p3/2 and 2p1/2 in TiO2, respectively. The strong peak at 530.1 eV in the O1s spectrum was the O–Ti bond in TiO2, while 531.0 eV and 532.7 eV were the surface-adsorbed water molecules, H2O and surface hydroxyl groups (-OH), respectively [39].
Figure S7 shows the UV–Vis diffuse reflectance absorption spectra of the composite material (350 °C, 450 °C, 550 °C, 650 °C, 750 °C). It can be seen from the figure that under different temperature conditions, the UV–visible light absorption properties of the composites were different. The B-TiO2/C3N4 prepared at the three temperatures of 350 °C, 450 °C, and 550 °C had better visible light absorption ability in both the ultraviolet and visible light regions, and the effect of 550 °C was the best. This result was caused by B element entering into the TiO2 lattice and g-C3N4 structure. According to the low electronegativity of B element, it can be inferred that the 2p orbital of B may be different from the 2p orbital of O. Hybridization was generated, thus, enabling enhanced visible light absorption [40]. Otherwise, the 650 °C and 750 °C had high light absorption capacity in the ultraviolet regions of 200–350 nm and 200–400 nm, respectively. The light absorption ability of the visible light region greater than 400 nm was weak. This may have been due to the gradual decomposition of the g-C3N4 structure in the composite material with the increase in temperature, resulting in a decrease in the visible light absorption capacity of the material. At the same time, although the temperature increased, it could also have caused the transformation of the TiO2 crystal phase, while the mixed-phase TiO2 had visible light absorption ability, but the effect was not obvious. Therefore, according to the UV–Vis diffuse reflectance spectrum analysis, it is shown that the B-TiO2/C3N4 composites prepared at 550 °C had good light absorption ability.
Based on the above characterization analysis description, we can roughly infer the preparation process of composite materials. The TiO2/C3N4 material was exfoliated by using the molten reaction environment provided by B2O3 during the heating process. At the same time, the gas was decomposed by the trace amount of g-C3N4, thereby forming a large number of loose pore structures. With the increase in calcination temperature, the generated polar -B=O bond replaced the H atom of the amino group on carbon nitride melon and formed an O–B–Ti or TiB2 structure, thereby finally forming a structurally stable B-TiO2/C3N4 material. The specific reaction process inference diagram is shown in Figure S8.

3.3. Photocatalytic Efficiencies

Figure 3 shows the dark adsorption and light-driven photocatalytic degradation ability of MB (a) and RhB (b) of B-TiO2/C3N4 composite materials (350 °C, 450 °C, 550 °C, 650 °C, and 750 °C). The specific experimental process was as follows: 30 mg of B-TiO2/C3N4 composite was added into 100 mL of MB solution with a concentration of 20 mg/L and 10 mg/L of RhB solution, respectively. A half-hour dark reaction time was set in the experiment and then the photocatalytic degradation experiment was carried out. The removal ability of RhB and MB was analyzed by catalytically degrading RhB and MB under the irradiation of a xenon light source with a wavelength greater than 400 nm.
The calculation of the removal efficiency of each pollutant by several catalyst materials was calculated by Formula (1). The experimental results of the dark reaction and the light reaction are shown in Figure 4a,b. The two figures show the adsorption–degradation curves of MB (20 mg/L) and RhB (10 mg/L) by the B-TiO2/C3N4 composite, respectively. The relationship among the adsorption–degradation efficiencies can be clearly seen from the figure: 550 °C > 650 °C > 450 °C > 750 °C > 350 °C. The B-TiO2/C3N4 prepared at the 550 °C calcination temperature had the best adsorption effect on the two dyes in the dark reaction adsorption process and the best removal efficiency (83.4% (MB), 64.1% (RhB)). At this temperature, B2O3 may undergo a complete molten state process, and become a good molten state reaction environment to fully contact with TiO2/C3N4. Therefore, the prepared B-TiO2/C3N4 had a larger specific surface area, exposing more adsorption sites and a reaction active site. Therefore, the doping of B element is beneficial to improve photocatalytic activity. When the temperature continued to rise above 600 °C, the g-C3N4 material began to decompose gradually, and the TiO2 gradually transformed from anatase type to rutile phase with lower photocatalytic reaction activity, so the photocatalytic reaction efficiency decreased gradually.
Meanwhile, multiple cycle experiments were carried out with MB and RhB. It can be seen from Figure 4 that after four cycles of repeated tests, the adsorption and removal rate of B-TiO2/C3N4 to 20 mg/L organic solution decreased from 83.4% to 70.1%(MB) and from 64.1% to 51.2% (RhB). The results showed that the adsorption effect of the material on MB and RhB was weakened gradually. This may have been because the pores on the surface of the material were damaged during the experiment. Meanwhile, the active sites were reduced due to the coverage of undegraded pollutants adsorbed on the surface of the material, resulting in a decrease in the adsorption and catalytic efficiency of the material after multiple experiments. Although the adsorption capacity of the material for the two pollutions was gradually weakened, the removal rate remained at around 95% (MB) and 80% (RhB) within two hours. This result shows that the B-TiO2/C3N4 composite material had better stability and repeatability.
The TOC in the photocatalytic degradation system reflected the mineralization degree of organic pollutants. The results of the TOC are shown in Figure S9. The adsorption process effectively reduced the value. With the progress of visible light, the TOC value experienced small decreases with the prolongation of the lighting time. This is because, during the degradation process, the pollutant molecules were decomposed from macromolecules to small molecules containing organic carbon, and the pollutants adsorbed on the surface of the material were also desorbed. The resorption–degradation process resulted in insignificant changes in the organic carbon content in the solution [41,42].

3.4. Adsorption Property

To demonstrate the adsorption property of B-TiO2/C3N4 (550 °C) composite materials, adsorption and removal rates of RhB and MB were chosen as the contaminant. The adsorption efficiency plays an important role in the rapid removal of pollutants.
The kinetic curve of adsorption of MB and RhB on B-TiO2/C3N4 is shown in Figure 5. From this figure, it can be seen that B-TiO2/C3N4 reached the adsorption equilibrium for the two pollutants within 30 min; the adsorption capacity of MB reached 56.25 mg/g; and the adsorption capacity of RhB reached 18.94 mg/g. Compared with other materials (as shown in Figure S1), the improvement of adsorption efficiency was due to the increase in specific surface area and the increase in adsorption sites, and the doping of B element in g-C3N4 expanded the large, π-bonded conjugated system, which was more conducive to the adsorption of MB [9].
The pseudo-first-order kinetic and pseudo-second-order kinetic models were fitted to the adsorption amounts of MB and RhB, respectively. The results are shown in Figure 6. The fitted correlation coefficients were: MB: R12 = 0.5017 (a), R22 = 0.9964 (b); RhB: R12 = 0.6872 (c), R22 = 0.9998 (d). According to the correlation coefficients fitted to the two pollutants, it can be seen that the pseudo-second-order kinetic model better fitted the adsorption behavior of B-TiO2/C3N4.
Figure 7 shows the fitting of the Langmuir and Freundlich adsorption isotherm models for the adsorption of MB (a) (b) and RhB (c) (d) on B-TiO2/C3N4, respectively. By comparing the correlation coefficients obtained by calculation, it was found that for the MB adsorption isotherm model, RL2 = 0.9946 was higher than that of the Freundlich model (RF2 = 0.6883), while the simulated adsorption isotherm model for RhB had the same or similar results (RL2: 0.9993 > RF2: 0.7395). This result showed that the adsorption of MB and RhB by B-TiO2/C3N4 was monolayer adsorption, and the adsorption was more favorable with the increase in pollutant concentration. Therefore, the adsorption of these two pollutants by B-TiO2/C3N4 was more inclined to the Langmuir adsorption isotherm model.

3.5. Analysis of Photocatalytic Mechanism

Based on the above experimental results, MB was selected as the target pollutant with the best removal effect, and the main substances involved in the photocatalytic reaction were determined by the free radical capture experiment. The experimental results are shown in Figure 8. The addition of EDTA-2Na had less of an effect on the degradation effect of MB, which indicates that the role of holes is less important in the catalytic reaction. After the addition of p-benzoquinone, the degradation of MB was inhibited, indicating that superoxide radical (∙O2−) is an important free radical involved in the photodegradation process. After the addition of tert-Butanol, the degradation of MB was also inhibited to a certain extent, so it can be concluded that hydroxyl radicals (∙OH) are also involved in the photocatalytic oxidation process. The above research results show that ∙O2− and ∙OH radicals were the active groups that mainly participated in the reaction during the degradation of MB by B-TiO2/C3N4.
Through mechanism analysis, it was found that ∙O2− and ∙OH were the main active groups involved in the photocatalytic reaction. Therefore, DMPO ESR was used to detect the ∙O2− and ∙OH produced by B-TiO2/C3N4 material under visible light above 400 nm. In addition, a comprehensive comparative analysis was conducted with other materials. It can be clearly seen from Figure 9a that after 10 min of light source irradiation, the order of ∙O2−ESR signal intensity was B-TiO2/C3N4 > TiO2/C3N4 > g-C3N4 > B-C3N4 > B-TiO2 > TiO2. The signal peak generated by B-TiO2/C3N4 was slightly higher than that of TiO2/C3N4 and significantly higher than that of the other three materials. In the ESR spectrum of ∙Oh, as shown in Figure 9b, the signal peak generated by B-TiO2/C3N4 was the strongest. According to the analysis results, the B-TiO2/C3N4 material produced the most ∙O2− and ∙OH in the visible light photocatalytic system, so it had a better photocatalytic effect.
Otherwise, the XPS characterization showed that the B element in B2O3 entered the TiO2 crystal structure. Therefore, in order to analyze the structural composition of B-TiO2 and B-C3N4, the conduction band relationship of the forbidden band width of B-TiO2 and B-C3N4 was analyzed. The specific results are shown in Figure 10a.
The photocatalytic reaction mechanism of B-TiO2/C3N4 composite material was deduced according to the radical trapping experiment. As shown in Figure 10b, under visible light conditions, the conduction band position of B-C3N4 was −0.88 eV, which was more negative than that of B-TiO2 (−0.31 eV) and E0 (O2/·O2− = −0.33 eV). The electrons could be captured by O2 to form ·O2− and finally formed H2O2 on the surface of the material. The holes were trapped by H2O and OH in the valence band of B-TiO2 to form hydroxyl radicals (·OH). These species are vital for photocatalytic degradation of organic pollution. Meanwhile, it can be determined that the B-TiO2/C3N4 composite material formed a Z-type heterostructure [43,44].

4. Conclusions

Herein, B-TiO2/C3N4 composites were prepared by modifying Z-type heterostructured TiO2/C3N4 composites using the molten reaction environment created by B2O3. The physical-chemical properties of B-TiO2/C3N4 composites prepared at different temperatures were analyzed by various characterization methods. Through characterization, it was found that the addition of B2O3 effectively increased the specific surface area of TiO2/C3N4 materials, thereby increasing the active sites for adsorption and photocatalysis. Crystal analysis showed that the modification of TiO2/C3N4 by B2O3 simultaneously doped B element into the TiO2 lattice and g-C3N4, which helped to inhibit the recombination of electron holes and improved the photocatalytic activity. At the same time, the doping of B element expanded the π-conjugated system of the g-C3N4 material, which is beneficial for the improvement of MB adsorption performance. Above all, this paper started with consideration of the two aspects of improving the adsorption capacity and photocatalytic oxidation capacity of the material and integrated the preparation, characterization, and adsorption–degradation mechanism analysis of the material, providing theoretical support and experimental basis for research on the synergistic removal of pollutants by adsorption and photocatalysis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijerph19148683/s1, Table S1: Comparison of MB removal rate of different materials, Figure S1: MB removal efficiency of several materials: TiO2, g-C3N4, B-TiO2, B-C3N4, TiO2/C3N4, B2-C3N4/TiO2 and B-TiO2/C3N4, Figure S2: MB removal efficiency of B-TiO2/C3N4 prepared with different doping amount, Figure S3: N2 adsorption-desorption curves of B-TiO2/C3N4 materials at different temperatures, Figure S4: XRD (X-ray diffraction) patterns of B2O3 and B-TiO2/C3N4-2 Series materials, Figure S5: FT-IR spectra for different temperature B-TiO2/C3N4, Figure S6: XPS spectra of B-TiO2/C3N4 material (a)B1s, (b)C1s, (c) N1s, (d) Ti2p and (e) O1s, Figure S7. The Diffuse reflectance spectra for different temperature B-TiO2/C3N4, Figure S8. Synthesis path of B-TiO2/C3N4 material, Figure S9. Changes of TOC in the experiment of adsorption-catalytic degradation of pollutants (a) MB (b) RhB. References [45,46,47] are cited in the supplementary materials.

Author Contributions

X.G. contributed substantially to the design of the study, the analyses, writing of the manuscript, and the original draft preparation. L.R. contributed to the project administration, funding acquisition, and verified the experimental data, replied to comments, and corrected the proof. Z.S. took responsibility for performing experiments and validation and made the figures. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by grants from the National Natural Science Foundation of China (no. 51775167) and the National Major Projects of Water Pollution Control and Management Technology (no. 2017ZX07204003).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ratshiedana, R.; Kuvarega, A.T.; Mishra, A.K. Titanium Dioxide and Graphitic Carbon Nitride–Based Nanocomposites and Nanofibres for the Degradation of Organic Pollutants in Water: A Review. Environ. Sci. Pollut. Res. 2021, 28, 10357–10374. [Google Scholar] [CrossRef] [PubMed]
  2. Xie, K.; Fang, J.; Li, L.; Deng, J.; Chen, F. Progress of Graphite Carbon Nitride with Different Dimensions in the Photocatalytic Degradation of Dyes: A Review. J. Alloys Compd. 2022, 901, 163589. [Google Scholar] [CrossRef]
  3. Hu, C.; Tu, S.; Tian, N.; Ma, T.; Zhang, Y.; Huang, H. Photocatalysis Enhanced by External Fields. Angew. Chem. Int. Ed. 2021, 60, 16309–16328. [Google Scholar] [CrossRef] [PubMed]
  4. Ahmaruzzaman, M.d.; Mishra, S.R. Photocatalytic Performance of g-C3N4 Based Nanocomposites for Effective Degradation/Removal of Dyes from Water and Wastewater. Mater. Res. Bull. 2021, 143, 111417. [Google Scholar] [CrossRef]
  5. Oh, J.; Shim, Y.; Lee, S.; Park, S.; Jang, D.; Shin, Y.; Ohn, S.; Kim, J.; Park, S. Structural Insights into Photocatalytic Performance of Carbon Nitrides for Degradation of Organic Pollutants. J. Solid State Chem. 2018, 258, 559–565. [Google Scholar] [CrossRef]
  6. Kuznetsov, V.N.; Serpone, N. Visible Light Absorption by Various Titanium Dioxide Specimens. J. Phys. Chem. B 2006, 110, 25203–25209. [Google Scholar] [CrossRef]
  7. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-Scheme Heterojunction Photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  8. Ejeta, S.Y.; Imae, T. Photodegradation of Pollutant Pesticide by Oxidized Graphitic Carbon Nitride Catalysts. J. Photochem. Photobiol. A Chem. 2021, 404, 112955. [Google Scholar] [CrossRef]
  9. Abe, M.; Kawashima, K.; Kozawa, K.; Sakai, H.; Kaneko, K. Amination of Activated Carbon and Adsorption Characteristics of Its Aminated Surface. Langmuir 2000, 16, 5059–5063. [Google Scholar] [CrossRef]
  10. Zhou, Z.; Li, K.; Deng, W.; Li, J.; Yan, Y.; Li, Y.; Quan, X.; Wang, T. Nitrogen Vacancy Mediated Exciton Dissociation in Carbon Nitride Nanosheets: Enhanced Hydroxyl Radicals Generation for Efficient Photocatalytic Degradation of Organic Pollutants. J. Hazard. Mater. 2020, 387, 122023. [Google Scholar] [CrossRef]
  11. Hossain, S.M.; Park, H.; Kang, H.-J.; Mun, J.S.; Tijing, L.; Rhee, I.; Kim, J.-H.; Jun, Y.-S.; Shon, H.K. Facile Synthesis and Characterization of Anatase TiO2/g-CN Composites for Enhanced Photoactivity under UV–Visible Spectrum. Chemosphere 2021, 262, 128004. [Google Scholar] [CrossRef] [PubMed]
  12. Serpone, N. Relative Photonic Efficiencies and Quantum Yields in Heterogeneous Photocatalysis. J. Adv. Oxid. Technol. 1997, 2, 203–216. [Google Scholar] [CrossRef]
  13. Yu, L.; Wang, Z.; Shi, L.; Yuan, S.; Zhao, Y.; Fang, J.; Deng, W. Photoelectrocatalytic Performance of TiO2 Nanoparticles Incorporated TiO2 Nanotube Arrays. Appl. Catal. B Environ. 2012, 113–114, 318–325. [Google Scholar] [CrossRef]
  14. Valero-Romero, M.J.; Santaclara, J.G.; Oar-Arteta, L.; van Koppen, L.; Osadchii, D.Y.; Gascon, J.; Kapteijn, F. Photocatalytic Properties of TiO2 and Fe-Doped TiO2 Prepared by Metal Organic Framework-Mediated Synthesis. Chem. Eng. J. 2019, 360, 75–88. [Google Scholar] [CrossRef]
  15. Kamegawa, T.; Shimizu, Y.; Yamashita, H. Superhydrophobic Surfaces with Photocatalytic Self-Cleaning Properties by Nanocomposite Coating of TiO2 and Polytetrafluoroethylene. Adv. Mater. 2012, 24, 3697–3700. [Google Scholar] [CrossRef]
  16. Horikoshi, S.; Serpone, N. Can the Photocatalyst TiO2 Be Incorporated into a Wastewater Treatment Method Background and Prospects. Catal. Today 2020, 340, 334–346. [Google Scholar] [CrossRef]
  17. Zhang, F.; Hong, B.; Zhao, W.; Yang, Y.; Bao, J.; Gao, C.; Sun, S. Ozone Modification as an Efficient Strategy for Promoting the Photocatalytic Effect of TiO2 for Air Purification. Chem. Commun. 2019, 55, 3757–3760. [Google Scholar] [CrossRef]
  18. Yuzer, B.; Aydın, M.I.; Con, A.H.; Inan, H.; Can, S.; Selcuk, H.; Kadmi, Y. Photocatalytic, Self-Cleaning and Antibacterial Properties of Cu(II) Doped TiO2. J. Environ. Manag. 2022, 302, 114023. [Google Scholar] [CrossRef]
  19. Hossain, S.M.; Tijing, L.; Suzuki, N.; Fujishima, A.; Kim, J.-H.; Shon, H.K. Visible Light Activation of Photocatalysts Formed from the Heterojunction of Sludge-Generated TiO2 and g-CN towards NO Removal. J. Hazard. Mater. 2022, 422, 126919. [Google Scholar] [CrossRef]
  20. Ojha, D.P.; Karki, H.P.; Kim, H.J. Design of Ternary Hybrid ATO/g-C3N4/TiO2 Nanocomposite for Visible-Light-Driven Photocatalysis. J. Ind. Eng. Chem. 2018, 61, 87–96. [Google Scholar] [CrossRef]
  21. Wang, P.; Guo, X.; Rao, L.; Wang, C.; Guo, Y.; Zhang, L. A Weak-Light-Responsive TiO2/g-C3N4 Composite Film: Photocatalytic Activity under Low-Intensity Light Irradiation. Environ. Sci. Pollut. Res. 2018, 25, 20206–20216. [Google Scholar] [CrossRef] [PubMed]
  22. Qumar, U.; Hassan, J.Z.; Bhatti, R.A.; Raza, A.; Nazir, G.; Nabgan, W.; Ikram, M. Photocatalysis vs Adsorption by Metal Oxide Nanoparticles. J. Mater. Sci. Technol. 2022. [Google Scholar] [CrossRef]
  23. Bahrudin, N.N.; Nawi, M.A.; Zainal, Z. Insight into the Synergistic Photocatalytic-Adsorptive Removal of Methyl Orange Dye Using TiO2/Chitosan Based Photocatalyst. Int. J. Biol. Macromol. 2020, 165, 2462–2474. [Google Scholar] [CrossRef] [PubMed]
  24. Zhu, Y.; Ji, H.; He, K.; Blaney, L.; Xu, T.; Zhao, D. Photocatalytic Degradation of GenX in Water Using a New Adsorptive Photocatalyst. Water Res. 2022, 220, 118650. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, M.; Zhang, D.; Han, J.; Liu, C.; Ding, Y.; Wang, Z.; Wang, A. Adsorption Enhanced Photocatalytic Degradation Sulfadiazine Antibiotic Using Porous Carbon Nitride Nanosheets with Carbon Vacancies. Chem. Eng. J. 2020, 382, 123017. [Google Scholar] [CrossRef]
  26. Guo, X.; Rao, L.; Wang, P.; Zhang, L.; Wang, Y. Synthesis of Porous Boron-Doped Carbon Nitride: Adsorption Capacity and Photo-Regeneration Properties. Int. J. Environ. Res. Public Health 2019, 16, 581. [Google Scholar] [CrossRef] [Green Version]
  27. Rodríguez-Narciso, S.; Lozano-Álvarez, J.A.; Salinas-Gutiérrez, R.; Castañeda-Leyva, N. A Stochastic Model for Adsorption Kinetics. Adsorpt. Sci. Technol. 2021, 2021, 5522581. [Google Scholar] [CrossRef]
  28. Zhu, Z.; Xie, J.; Zhang, M.; Zhou, Q.; Liu, F. Insight into the Adsorption of PPCPs by Porous Adsorbents: Effect of the Properties of Adsorbents and Adsorbates. Environ. Pollut. 2016, 214, 524–531. [Google Scholar] [CrossRef]
  29. She, X.; Liu, L.; Ji, H.; Mo, Z.; Li, Y.; Huang, L.; Du, D.; Xu, H.; Li, H. Template-Free Synthesis of 2D Porous Ultrathin Nonmetal-Doped g-C3N4 Nanosheets with Highly Efficient Photocatalytic H2 Evolution from Water under Visible Light. Appl. Catal. B Environ. 2016, 187, 144–153. [Google Scholar] [CrossRef]
  30. Guo, Q.; Ma, Z.; Zhou, C.; Ren, Z.; Yang, X. Single Molecule Photocatalysis on TiO2 Surfaces. Chem. Rev. 2019, 119, 11020–11041. [Google Scholar] [CrossRef]
  31. Li, L.; Hong, W.B.; Yan, X.J.; Chen, X.M. Preparation and Microwave Dielectric Properties of B2O3 Bulk. Int. J. Appl. Ceram. Technol. 2019, 16, 2047–2052. [Google Scholar] [CrossRef]
  32. Photocatalytic Activity Enhancement of Core-Shell Structure g-C3N4@TiO2 via Controlled Ultrathin g-C3N4 Layer-ScienceDirect. Available online: https://www.sciencedirect.com/science/article/abs/pii/S0926337317307403 (accessed on 20 June 2022).
  33. Mozzi, R.L.; Warren, B.E. The Structure of Vitreous Boron Oxide. J. Appl. Cryst. 1970, 3, 251–257. [Google Scholar] [CrossRef]
  34. Danon, A.; Stair, P.C.; Weitz, E. FTIR Study of CO2 Adsorption on Amine-Grafted SBA-15: Elucidation of Adsorbed Species. J. Phys. Chem. C 2011, 115, 11540–11549. [Google Scholar] [CrossRef]
  35. Quiñones, D.H.; Rey, A.; Álvarez, P.M.; Beltrán, F.J.; Li Puma, G. Boron Doped TiO2 Catalysts for Photocatalytic Ozonation of Aqueous Mixtures of Common Pesticides: Diuron, o-Phenylphenol, MCPA and Terbuthylazine. Appl. Catal. B Environ. 2015, 178, 74–81. [Google Scholar] [CrossRef] [Green Version]
  36. Ahmed, T.; Ammar, M.; Saleem, A.; Zhang, H.; Xu, H. Z-Scheme 2D-m-BiVO4 Networks Decorated by a g-CN Nanosheet Heterostructured Photocatalyst with an Excellent Response to Visible Light. RSC Adv. 2020, 10, 3192–3202. [Google Scholar] [CrossRef] [Green Version]
  37. Wei, H.; McMaster, W.A.; Tan, J.Z.Y.; Cao, L.; Chen, D.; Caruso, R.A. Mesoporous TiO2/g-C3N4 Microspheres with Enhanced Visible-Light Photocatalytic Activity. J. Phys. Chem. C 2017, 121, 22114–22122. [Google Scholar] [CrossRef]
  38. Guo, Y.; Wang, R.; Wang, P.; Rao, L.; Wang, C. Developing a Novel Layered Boron Nitride–Carbon Nitride Composite with High Efficiency and Selectivity To Remove Protonated Dyes from Water. ACS Sustain. Chem. Eng. 2019, 7, 5727–5741. [Google Scholar] [CrossRef]
  39. Zhu, J.; Chen, F.; Zhang, J.; Chen, H.; Anpo, M. Fe3+-TiO2 Photocatalysts Prepared by Combining Sol–Gel Method with Hydrothermal Treatment and Their Characterization. J. Photochem. Photobiol. A Chem. 2006, 180, 196–204. [Google Scholar] [CrossRef]
  40. Barkul, R.P.; Sutar, R.S.; Patil, M.K.; Delekar, S.D. Photocatalytic Degradation of Organic Pollutants by Using Nanocrystalline Boron-Doped TiO2 Catalysts. ChemistrySelect 2021, 6, 3360–3369. [Google Scholar] [CrossRef]
  41. Beltrán, F.J.; Aguinaco, A.; García-Araya, J.F. Kinetic Modelling of TOC Removal in the Photocatalytic Ozonation of Diclofenac Aqueous Solutions. Appl. Catal. B Environ. 2010, 100, 289–298. [Google Scholar] [CrossRef]
  42. Lee, J.C.; Park, S.; Park, H.-J.; Lee, J.-H.; Kim, H.-S.; Chung, Y.-J. Photocatalytic Degradation of TOC from Aqueous Phenol Solution Using Solution Combusted ZnO Nanopowders. J. Electroceram. 2009, 22, 110–113. [Google Scholar] [CrossRef]
  43. Tatykayev, B.; Chouchene, B.; Balan, L.; Gries, T.; Medjahdi, G.; Girot, E.; Uralbekov, B.; Schneider, R. Heterostructured g-CN/TiO2 Photocatalysts Prepared by Thermolysis of g-CN/MIL-125(Ti) Composites for Efficient Pollutant Degradation and Hydrogen Production. Nanomaterials 2020, 10, 1387. [Google Scholar] [CrossRef] [PubMed]
  44. Karpuraranjith, M.; Chen, Y.; Rajaboopathi, S.; Ramadoss, M.; Srinivas, K.; Yang, D.; Wang, B. Three-Dimensional Porous MoS2 Nanobox Embedded g-C3N4@TiO2 Architecture for Highly Efficient Photocatalytic Degradation of Organic Pollutant. J. Colloid Interface Sci. 2022, 605, 613–623. [Google Scholar] [CrossRef] [PubMed]
  45. Feng, J.; Chen, T.; Liu, S.; Zhou, Q.; Ren, Y.; Lv, Y.; Fan, Z. Improvement of g-C3N4 Photocatalytic Properties Using the Hummers Method. J. Colloid Interface Sci. 2016, 479, 1–6. [Google Scholar] [CrossRef]
  46. Kwon, D.; Kim, J. Silver-Doped ZnO for Photocatalytic Degradation of Methylene Blue. Korean J. Chem. Eng. 2020, 37, 1226–1232. [Google Scholar] [CrossRef]
  47. Mo, Z.; She, X.; Li, Y.; Liu, L.; Huang, L.; Chen, Z.; Zhang, Q.; Xu, H.; Li, H. Synthesis of g-C3N4 at Different Temperatures for Superior Visible/UV Photocatalytic Performance and Photoelectrochemical Sensing of MB Solution. RSC Adv. 2015, 5, 101552–101562. [Google Scholar] [CrossRef]
Figure 1. SEM of B-TiO2/C3N4 under different calcination conditions: (a) 350 °C; (b) 450 °C; (c) 550 °C; (d) 650 °C; (e) 750 °C.
Figure 1. SEM of B-TiO2/C3N4 under different calcination conditions: (a) 350 °C; (b) 450 °C; (c) 550 °C; (d) 650 °C; (e) 750 °C.
Ijerph 19 08683 g001
Figure 2. SEM (a,b) and TEM (c,d) spectra of B-TiO2/C3N4 (550 °C) composite materials.
Figure 2. SEM (a,b) and TEM (c,d) spectra of B-TiO2/C3N4 (550 °C) composite materials.
Ijerph 19 08683 g002
Figure 3. The adsorption–photocatalytic degradation curves of MB (a) and RhB (b) over B-TiO2/C3N4.
Figure 3. The adsorption–photocatalytic degradation curves of MB (a) and RhB (b) over B-TiO2/C3N4.
Ijerph 19 08683 g003
Figure 4. Cycling experiments of B-TiO2/C3N4 for the adsorption–photocatalytic degradation of MB (a) and RhB (b).
Figure 4. Cycling experiments of B-TiO2/C3N4 for the adsorption–photocatalytic degradation of MB (a) and RhB (b).
Ijerph 19 08683 g004
Figure 5. Adsorption kinetics curves of B-TiO2/C4 on MB and RhB.
Figure 5. Adsorption kinetics curves of B-TiO2/C4 on MB and RhB.
Ijerph 19 08683 g005
Figure 6. (a) Plots of the pseudo-first-order model (MB); (b) Plots of the pseudo-second-order model (MB); (c) Plots of the pseudo-first-order model (RhB); (d) Plots of the pseudo-second-order model (RhB).
Figure 6. (a) Plots of the pseudo-first-order model (MB); (b) Plots of the pseudo-second-order model (MB); (c) Plots of the pseudo-first-order model (RhB); (d) Plots of the pseudo-second-order model (RhB).
Ijerph 19 08683 g006
Figure 7. The linearized of Langmuir adsorption isotherm (a) MB (c) RhB and Freundlich adsorption isotherm (b) MB (d) RhB adsorption by B-TiO2/C3N4.
Figure 7. The linearized of Langmuir adsorption isotherm (a) MB (c) RhB and Freundlich adsorption isotherm (b) MB (d) RhB adsorption by B-TiO2/C3N4.
Ijerph 19 08683 g007
Figure 8. The capture experiments of MB by B-TiO2/C3N4.
Figure 8. The capture experiments of MB by B-TiO2/C3N4.
Ijerph 19 08683 g008
Figure 9. Spectrum of DMPO spin trapping ESR of different materials (a) DMPO-O2 and (b) DMPO-·OH.
Figure 9. Spectrum of DMPO spin trapping ESR of different materials (a) DMPO-O2 and (b) DMPO-·OH.
Ijerph 19 08683 g009
Figure 10. The photocatalytic reaction mechanism of B-TiO2/C3N4: (a) The VBXPS of B-TiO2 and B-CN; (b) Scheme of proposed mechanism for degradation of B-TiO2/C3N4.
Figure 10. The photocatalytic reaction mechanism of B-TiO2/C3N4: (a) The VBXPS of B-TiO2 and B-CN; (b) Scheme of proposed mechanism for degradation of B-TiO2/C3N4.
Ijerph 19 08683 g010
Table 1. BET surface area, average pore size, and pore volume of B-TiO2/C3N4 material at different temperatures.
Table 1. BET surface area, average pore size, and pore volume of B-TiO2/C3N4 material at different temperatures.
Sample350 °C450 °C550 °C650 °C750 °C
BET (m2/g)13.38113.47258.65434.62225.795
Average pore size (nm)6.3117.8096.96012.73314.263
Average pore volume (cm3/g)0.0900.0730.1300.1620.103
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, X.; Rao, L.; Shi, Z. Preparation of High-Porosity B-TiO2/C3N4 Composite Materials: Adsorption–Degradation Capacity and Photo-Regeneration Properties. Int. J. Environ. Res. Public Health 2022, 19, 8683. https://doi.org/10.3390/ijerph19148683

AMA Style

Guo X, Rao L, Shi Z. Preparation of High-Porosity B-TiO2/C3N4 Composite Materials: Adsorption–Degradation Capacity and Photo-Regeneration Properties. International Journal of Environmental Research and Public Health. 2022; 19(14):8683. https://doi.org/10.3390/ijerph19148683

Chicago/Turabian Style

Guo, Xiang, Lei Rao, and Zhenyu Shi. 2022. "Preparation of High-Porosity B-TiO2/C3N4 Composite Materials: Adsorption–Degradation Capacity and Photo-Regeneration Properties" International Journal of Environmental Research and Public Health 19, no. 14: 8683. https://doi.org/10.3390/ijerph19148683

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