Next Article in Journal
New Process Combining Fe-Based Chemical Looping and Biomass Pyrolysis for Cogeneration of Hydrogen, Biochar, Bio-Oil and Electricity with In-Suit CO2 Separation
Previous Article in Journal
Influence of Waste Toner on Asphalt Binder: Chemical and Rheological Characterization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Multisystem Photocatalytic Degradation of Anionic and Cationic Dyes Using Graphitic Phase Carbon Nitride

1
College of Environmental and Energy Engineering, Anhui Jianzhu University, Hefei 230601, China
2
Anhui Key Laboratory of Environmental Pollution Control and Waste Resource Utilization, Anhui Jianzhu University, Hefei 230601, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(6), 2796; https://doi.org/10.3390/molecules28062796
Submission received: 16 February 2023 / Revised: 7 March 2023 / Accepted: 9 March 2023 / Published: 20 March 2023

Abstract

:
Graphitic phase carbon nitride (g-C3N4) is a promising photocatalytic environmental material. For this study, the graphitic phase carbon nitride was prepared using a thermal polymerization method. The characteristic peaks, structures, and morphologies were determined using Fourier-transform infrared spectroscopy (FT-IR), X-ray diffractometry (XRD), and scanning electron microscopy (SEM), respectively. Under the synergetic visible light catalysis of H2O2 and Na2S2O8, the degradation effects of g-C3N4 on the anionic dye methyl orange (MO) and the cationic dye rhodamine b (Rhb) were investigated. The effects of adding different volumes of H2O2 and Na2S2O8 were likewise tested. The results showed that the above two synergistic systems increased the degradation rates of MO and Rhb by 2.5 and 3.5 times, respectively, compared with pure g-C3N4, and that the degradation rates of both MO and Rhb reached 100% within 120 min and 90 min, respectively, in accordance with the primary reaction kinetics. When H2O2 and Na2S2O8 were added dropwise at 10 mL each, the degradation rates of MO and Rhb were 82.22% and 99.81%, respectively, after 30 min of open light. The results of experiments upon both zeta potential and radical quenching showed that ·OH and ·O2 were the main active radicals for dye degradation in our synergistic system. In addition, stability tests showed that the photocatalysts in the synergistic system still had good reusability. Therefore, the use of a synergistic system can effectively reduce the photogenerated electron-hole pair complexation rate, representing a significant improvement in both photocatalytic degradation and for stability levels.

1. Introduction

With the rapid developments in industrialization, water pollution has become one of the main problems affecting people’s everyday lives [1,2,3,4]. Among them, dye wastewater is an important part of water pollution [5,6,7]. Dyes are widely used in various industries such as textiles, leather, and paper, and it has been reported that about 30% of synthetic dyes are discharged into the environment [8,9,10]. Dye wastewater is complex and difficult to degrade, contains chromogenic and polar groups, and can cause eutrophication of water bodies [11,12,13]. Dyes can induce allergies and cancer, damage the respiratory system, nervous system and skin, and cause harm to the human body. Therefore, finding an effective and environmentally friendly method to degrade dyestuffs has become an urgent problem [14].
Photocatalytic methods, being one of the more advanced oxidation techniques, can oxidize pollutants into harmless small molecules, such as H2O and CO2, using free radicals generated during their reaction with light [15,16,17]. Photocatalytic technology can be applied to a wide range of pollutants with a wide range of situations. Lv et al. [18] used photocatalytic technology to degrade common air pollutants, with removal rates above 80%. Zeng et al. [19] used photocatalytic technology to treat low concentration ammonia nitrogen wastewater, with removal rates above 90%. TiO2 photocatalytic materials were discovered early on, but they have a large band gap in energy, and can only produce carrier electron-hole pairs in the UV range, its solar energy utilization being only about 4% [20,21]. Similarly, ZnO has a relatively wide band gap value (3.37 eV) and has a photocatalytic response only under UV light [22]. People have been looking for photocatalytic materials that can accept visible light. In recent years, graphitic phase carbon nitride (g-C3N4), with its two-dimensional layered structure, has been widely used in photocatalysis because of the narrow width of its forbidden band (about 2.7 eV), its high chemical stability, non-toxicity, and low preparation cost [23,24,25,26,27,28]. Saeed et al. [29] used photocatalysis as an effective tool to study dye degradation, and Zhou et al. [30] prepared g-C3N4-based photocatalytic concrete, which degraded 80% of their methylene blue within 30 min. Pandey et al. [31] removed 96% and 93% of PNP and Rhb dyes, respectively, using photocatalysis.
Dyes can be classified into cationic dyes, anionic dyes, and nonionic dyes [32]. The more common cationic dyes include rhodamine b (Rhb), malachite green, and methylene blue (MB), and common anionic dyes include methyl orange (MO) and Congo red [33,34]. Ivanenko et al. [35] achieved a good photocatalytic effect by degrading anionic and cationic dyes with ZnO synthesized using precipitation, but this method uses only 4% of the available sunlight in using only UV light, which is not a good use of sunlight and is therefore of no practical use. Gupta et al. [36] doped ZnSe with Cu under visible light conditions and cooperated with H2O2 to improve the degradation of methylene blue, where 15% Cu doping was shown to be able to achieve 98.09% MB degradation in 180 min; however, this photocatalytic material is complicated to prepare, and has some dangerous qualities and a long degradation time. The photocatalyst used in this study is simple to prepare, has high yield, and can degrade 100% of Rhb in 90 min and 100% of MO in 150 min using a synergistic reaction. At present, there are comparatively few studies regarding the use of g-C3N4 for the photocatalytic degradation of cationic and anionic dyes, and synergistic photocatalysis can greatly improve the reaction rate and degradation rate based on dyes in materials. Lu et al. [37] have achieved better results in the synergistic amine photocatalytic synthesis of imines using carbon-coated Pd/TiO2, which can exhibit excellent imine selectivity. Under visible light irradiation, H2O2 can be excited using light energy or electrons to produce superoxide radicals (OH), and Na2S2O8 can be excited using electrons to produce sulfate radicals (SO4·) [38,39]. Thus, hydrogen peroxide (H2O2) and sodium persulfate (Na2S2O8) can be used for synergistic photocatalytic reactions. Su et al. [40] treated 2,4-dinitroanisole (DNAN) with ultraviolet (UV) photolysis combined with hydrogen peroxide (H2O2) oxidation. H2O2 can be adsorbed on active sites and then activate these sites. Liu et al. [41] used H2O2 as a green oxidant because its by-product is H2O. Methanol was then oxidized into methylformate (MF) using H2O2 oxidation.
In this study, g-C3N4 was prepared with a thermal polymerization method of simulating sunlight, and its characteristic peaks, structure, and morphology were characterized using Fourier-transform infrared spectroscopy (FT-IR), an X-ray diffractometer (XRD), and scanning electron microscopy (SEM). The photocatalytic effects of g-C3N4 on the cationic dye Rhb and the anionic dye MO were also investigated in concert with H2O2 and Na2S2O8 oxidants, both in accordance with the primary reaction kinetics. The degradation mechanism of the different systems was inferred from the quenching experiments of free radicals and the detection of zeta potential, and it was found that the main free radicals of the anionic dyes in the different synergistic systems were exactly opposite, which might be related to the acidity and basicity of the dyes themselves, which are yet to be explored.

2. Results and Discussion

2.1. Structural Characterization

2.1.1. XRD and FTIR Analysis

Figure 1a shows the FT-IR spectrum of g-C3N4. The surface functional groups of the prepared materials have been highlighted. The absorption peak at 809 cm−1 is caused by the deformation vibration of the N-C=N bond in the triazine ring, the absorption peak within the wave number from 1245 to 1631 cm−1 is due to the stretching vibration of the C-N bond, and the broad peak at 3000–3500 cm−1 is due to the stretching vibration of the -OH group on the surface of g-C3N4 [42,43,44,45].
As shown in Figure 1b, the crystal structure of g-C3N4 was determined using XRD, from which the characteristic peaks of the sample at 13.10 °C and 27.55 °C can be seen, where 13.10 °C and 27.55 °C correspond to the (100) and (002) crystal planes, respectively. The first diffraction peak is due to stacking within the triazine structural layer of the repeating unit in the plane, and the second diffraction peak is due to the interlayer stacking of the conjugated aromatic system that results from this [46,47]. This is a characteristic peak unique to the graphite structure, indicating the successful preparation of g-C3N4 [48,49].

2.1.2. SEM Analysis

Figure 2 shows the SEM image of g-C3N4, showing its change in morphology from layered stacking into an irregular, porous, tubular shape; this coral-like fluffy pore structure increases the specific surface area of the photocatalyst and provides more active sites for degrading pollutants, which can improve the degradation rate of photocatalysis to some extent [50,51,52,53].

2.2. Evaluation of the Degradation Performance of Dyes

In order to study the degradation performance of cationic and anionic dyes under two systems of H2O2 and Na2S2O8, g-C3N4 was added to 150 mL Rhb and 100 mL MO aqueous solutions, respectively, and then a certain amount of H2O2 and Na2S2O8 was added dropwise. Control experiments were conducted under LED, and the results are shown in Figure 3 and Figure 4.
As shown in Figure 3a, the concentration of Rhb solution decreased to a certain extent within 60 min of reaction with darkness, which was due to the adsorption of g-C3N4 itself. The degradation rate from pure g-C3N4 on the Rhb solution was 45.66% when the adsorption-desorption equilibrium was reached after direct light exposure for 90 min. In contrast, the removal effect of Rhb was significantly improved by adding H2O2 and Na2S2O8, respectively, and both reached a 100% degradation rate, a result which was 2.19 times higher than that of using pure g-C3N4 (see Table 1). This indicates that each of the two synergistic systems, H2O2 and Na2S2O8, played an important role in the photocatalytic degradation process of Rhb solution using g-C3N4, both in terms of reaction rate and degradation rate. Among them, the Na2S2O8 synergistic system showed somewhat higher speeds of degradation in the Rhb solution. In order to better observe the degradation pattern, the first-order reaction kinetics were fitted with experimental data according to Equation (11), and the results are shown in Figure 3b. The degradation rate of pure g-C3N4 was the slowest, at 0.02207 min−1; the degradation rate of H2O2 synergistic system was 0.06907 min−1; the highest degradation rate, that of the Na2S2O8 synergistic system, was 0.10511 min−1.
As shown in Figure 4a, the degradation rate of MO solution reached 15.18% after 60 min of adsorption-desorption, which is likewise due to the adsorption of g-C3N4 itself. After 90 min of exposure to open light, the degradation rates of the pure g-C3N4, H2O2, and Na2S2O8 synergistic systems reached 43.26%, 99.90%, and 98.70%, respectively (see Table 2). It is obvious that both g-C3N4 synergistic H2O2 and Na2S2O8 systems had a great degradation ability for MO solutions; among them all, g-C3N4 synergistic H2O2 system had higher degradation rate and reaction rate. As shown in Figure 4b, the MO photocatalytic process was consistent with quasi primary kinetics, and the degradation rates of pure g-C3N4, H2O2, and Na2S2O8 synergistic systems were k1 = 0.00563 min−1, k2 = 0.07753 min−1, and k3 = 0.03978 min−1, respectively.
In order to observe the degradation of dyes in each synergistic system, the absorption spectra of Rhb and MO under different catalytic conditions were measured using an UV spectrophotometer at various times during the experiment, with the results shown in Figure 5. Under the synergistic systems, Rhb was completely degraded within 90 min, while MO was completely degraded within 150 min and 210 min, respectively. After reviewing the literature summaries for comparison, it was found that the photocatalyst in this study is comparatively simple to prepare and can be mass produced [54,55,56,57,58,59]. With synergistic photocatalysis, 100% of the dye can be degraded in a relatively short time, as shown in Table 3.

2.3. Effect of Volume Content

2.3.1. Effect of Na2S2O8 Levels on Rhb Degradation

The g-C3N4 synergistic Na2S2O8 system has higher speeds for Rhb solutions, so the photocatalytic degradation of the Rhb solution across different volumes of Na2S2O8 was investigated. Different volumes of Na2S2O8 were added dropwise to each solution at the end of the dark reaction, and the fixed light time across all those exposed to light was 30 min, with no light being used in the control experiment. The results of this are shown in Figure 6.
In the absence of light, the degradation rate increased with the increase in the amount of Na2S2O8 added dropwise. As can be seen from Figure 6, the Rhb degradation rate was only 7% without the addition of Na2S2O8; when 10 mL Na2S2O8 was added dropwise, the Rhb degradation rate increased by about 53% compared with the solution without Na2S2O8.
In the presence of light, the larger the volume of Na2S2O8, the better the synergistic photocatalytic effect. When Na2S2O8 was not added, the degradation rate was 17%. When the dropwise amount was greater than 0.5 mL, it began to increase rapidly, and the degradation rate increased from 22% to 80%, indicating that when Na2S2O8 was added dropwise to a certain amount, it also greatly promoted the redox reactions. When 10 mL of Na2S2O8 was added dropwise, the Rhb was basically completely degraded, with a degradation rate of 100%, which was 40% higher at this time compared with not having been exposed to light.

2.3.2. Effect of H2O2 Volume on MO Degradation

For the MO solution, H2O2 synergistic system has better photocatalytic effect, so it is important to explore the MO degradation rate under different H2O2 volume conditions. Different amounts of H2O2 were added dropwise at the end of the dark reaction for 60 min, respectively, and the light time was fixed to be 30 min for both. The light time was changed for the dark reaction, to serve as a control. The results are shown in Figure 7.
In the absence of light, the MO degradation rate was 15% when H2O2 was not added dropwise. However, the MO degradation rate decreased when the volume was increased from 0.1 mL to 0.2 mL (degradation rate changed from 18% to 17%), which was due to the H2O2 catalytically generated during the addition of 0.1 mL; the degradation rate increased starting from the increase to 2 mL (with a resulting degradation rate of 27%).
The degradation rates were all significantly increased when there was light, but when the amount of H2O2 added dropwise was 0.1 mL, the degradation of MO solution was inhibited to some extent. It is possible that a small amount of H2O2 was adsorbed onto the surface of g-C3N4 and occupied a certain spot. When 0.2 mL of H2O2 was added, the degradation rate of MO solution gradually increased (degradation rate of 32%); when 10 mL of H2O2 was added dropwise, the degradation rate of MO was 82%.

2.4. Investigation of the Mechanism of Different Systems

There are many studies on the photocatalytic degradation of Rhb and MO solutions using g-C3N4. After turning on the light source, g-C3N4 is induced to produce electrons (e) and holes (h+) as in Equation (1). H2O is then oxidized with holes to produce ·OH and O2, as shown in Equations (2) and (3). O2 will be reduced with electrons to produce ·O2, as in Equation (4). The generated reactive radicals (·OH, ·O2) and cavities will degrade Rhb and MO solutions, as shown in Equation (5) [60,61,62].
g-C 3 N 4 + hv     g-C 3 N 4   ( e + h + )
H 2 O + h +     · OH + H +
6 H 2 O + 4 h +   4 H + + O 2  
O 2 + 2 e     · O 2          
h + / · O 2 / · OH + Rhb / MO     degradation   products    

2.4.1. Detection of Zeta Potential

As shown in Table 4, the mechanism of the two synergistic systems was further explored by measuring their zeta potential. g-C3N4 had an initial potential of −14.4 mV, indicating that the surfaces of its particles were mainly negatively charged. When the dye was then added, the negative surface charge appeared to increase in different cases, due to photocatalysis. The potentials of Rhb and MO were −27 mV and −19.6 mV, respectively. In contrast, when H2O2 was added, the zeta potential of Rhb and MO decreased to −20.3 mV and –12.6 mV, respectively. This is due to the fact that H2O2 also acts as an electron trapping agent in the synergistic systems, resulting in a decrease in negative surface charge.
Therefore, it can be presumed that H2O2 will undergo a reduction reaction with e- to form ·OH, as in Equation (6). When Na2S2O8 is added, the zeta potential did not change much, probably because the reaction of Equation (7) does not cause a change in charge [63,64]. As in Equation (8), SO4· can also convert Rhb and MO into small molecules. Under the energy excitation of visible light, H2O and O2 will produce H2O2, and H2O2 will produce OH (see Equations (9) and (10)).
H 2 O 2 + e     · OH + OH
H + + S 2 O 8 2     2 SO 4 · + H +  
SO 4 · + Rhb / MO     degradation   products
O 2 + 2 H 2 O + hv     H 2 O 2
H 2 O 2 + hv     · OH
From the above reactions, it can be seen that holes (h+), superoxide radicals (·O2), hydroxyl radicals (·OH), and sulfate radicals (SO4·) can degrade rhodamine b and methyl orange solutions.

2.4.2. Free Radical Capture Experiments

In order to determine the main active substances in the photocatalytic degradation of the cationic dye Rhb and anionic dye MO within different synergistic systems, quenching experiments of holes (h+) and radicals were carried out, and a blank control was made (see Table 5). Methanol (CH3OH), p-benzoquinone (C6H4O2), diphenylamine (C12H11N), and ethanol (CH3CH2OH) were chosen as quenching agents for h+, ·O2, ·OH, and SO4·, respectively [65,66]. The quenching reaction times were 90 min for all the Rhb solutions. The time for the quenching reaction of the MO solution was 150 min for the H2O2 synergistic system and 210 min for the Na2S2O8 synergistic system.
As can be seen in Figure 8a, the degradation of Rhb was 100% in the absence of any quenching agent, and the addition of methanol had almost no effect on the degradation of Rhb solution (showing an inhibition rate of 0.25%). The addition of both p-benzoquinone and diphenylamine had an inhibitory effect on the reaction (the inhibition rates were 30.15% and 64.6%, respectively). The degradation rate decreased from 100% to 35.4% with the addition of diphenylamine, which was 34.45% higher than that of p-benzoquinone, indicating that ·OH was the main active substance within this synergistic system. In Figure 8b, the addition of methanol hardly affected the degradation rate of Rhb. The addition of p-benzoquinone, diphenylamine, and ethanol all significantly inhibited the degradation of Rhb. The inhibition rate of diphenylamine was 4.88%, yet p-benzoquinone had the most prominent effect with 33.88% inhibition, which is about twice as much as ethanol (with an inhibition rate of 16.47%), indicating that ·O2 is the main reactive radical.
In Figure 9a, the addition of methanol inhibited 8.59% of the MO degradation. The degradation of MO decreased from 100% to 20.08%, and 41.26% after the addition of p-benzoquinone and diphenylamine, respectively, indicating that ·O2 and ·OH are the main active substances in the H2O2 synergistic g-C3N4 system. The inhibition rate of benzoquinone (79.92%) was 21.18% higher than that of diphenylamine (an inhibition rate of 58.74%), such that ·O2 was the more dominant reactive radical in this system. In Figure 9b, the addition of methanol and ethanol did not have much effect on the degradation of MO (both inhibition rates were 0). In addition, the degradation of MO was 48.65% and 66.73% after quenching with ·O2 and ·OH, respectively. ·OH was the more dominant reactive radical in this system.
As suggested by the above results, the main reactive substances of anionic and cationic dyes in different synergistic systems are exactly opposite (see Table 6). For the cationic dye Rhb, ·OH and ·O2 are the most dominant reactive substances in the H2O2 and Na2S2O8 synergistic g-C3N4 photocatalytic degradation systems, respectively. In contrast, in the anionic dye MO, ·O2 was the most dominant reactive radical in the H2O2 synergistic system, and ·OH was the most dominant reactive substance in the Na2S2O8 synergistic system. Thus, it is speculated that this phenomenon is related to contrasting pH level between the anionic and cationic dyes, which are acidic and alkaline, respectively. The mechanism for this is to be further investigated.

2.4.3. Stability Testing

The stability and recyclability of the photocatalyst are also important properties for its practical application. In order to determine the stability of the synergistic photocatalytic reaction, repeated experiments were carried out, as shown in Figure 10a. After five cycles, the degradation rate of Rhb in the g-C3N4/H2O2 synergistic system only showed a very slight activity decrease, remaining above 98%. By contrast, the MO degradation rate decreased, and the degradation rate was 82% in the 5th cycle, but the first four cycles were above 93%. As shown in Figure 10b, the degradation rates of Rhb and MO in the g-C3N4/Na2S2O8 synergistic system did not change much and remained above 90% after five cycles. To further investigate the structure and chemical stability of the photocatalyst, FTIR analysis was performed on the samples before and after the photocatalytic reaction. As shown in Figure 11, the FTIR spectra before and after the reaction were almost the same. This indicates that the g-C3N4 in the synergistic system not only has good stability during the degradation of Rhb and MO, but also has a good reusability performance. In addition, the slight decrease in photocatalytic activity is caused by the mass loss of the photocatalyst during the cycling process.

2.4.4. Principle of Photocatalysis

As shown in Figure 12, based on the preliminary experimental tests and theoretical analysis, we hypothesized the mechanism of g-C3N4 synergistic multisystem photocatalytic degradation of anionic and cationic dyes (Rhb and MO). Under the irradiation of visible light, the electrons on the valence band (VB) of g-C3N4 will be excited to the conduction band (CB), producing photogenerated electrons and holes, i.e., e and h+. The presence of two oxidants, Na2S2O8 and H2O2, will reduce the complex rate of e- and h+, so that more photogenerated carriers will be involved in the degradation of Rhb and MO. h+, ·O2, ·OH, SO4· can all degrade rhodamine b and methyl orange solutions. However, in synergistic systems, the main reactive radicals for the degradation of anionic dyes are ·OH and ·O2. The reduction reaction of O2 and H2O2 produces ·OH and ·O2; the oxidation reaction of H2O produces ·OH, and then the anionic dyes (Rhb and MO) are degraded to small molecules by the ·OH and ·O2.

3. Experimental Part

3.1. Reagents and Apparatus

Experimental reagents: urea, potassium bromide (Tianjin Damao Chemical Reagent Factory, Tianjin, China); rhodamine b (Rhb) (Tianjin Damao Chemical Reagent Factory, Tianjin, China); methyl orange (MO) (Tianjin Guangfu Technology Development Co., Ltd. Tianjin, China); 30% hydrogen peroxide (30% H2O2), and sodium persulfate (Na2S2O8) (Sinopharm Chemical Reagent Co., Ltd. Shanghai, China); p-benzoquinone (Shanghai Maclean Biochemical Technology Co. Ltd. Shanghai, China); anhydrous ethanol (Tianjin Zhiyuan Chemical Reagent Co., Ltd. Tianjin, China). All the above drugs were analytically pure (AR), and the experimental water was deionized water.
Experimental instruments: optical dark box (GXAS345); ultrasonic cleaner (KQ-100B); muffle furnace and UV–Vis spectrophotometer (UV-26001); Nicolet 380 FTIR spectrometer; Regulus 8100 scanning electron microscope (SEM); X-ray photoelectron spectrometer (XRD).

3.2. Preparation of Catalyst

Weigh a certain amount of urea, wrap it with tinfoil, place it in a muffle furnace and heat it up to 520 °C, hold it for 2 h, then heat it up to 550 °C and hold it for 2 h, wait for it to cool down to room temperature, transfer the roasted light yellow solid to a mortar and grind it to get g-C3N4.

3.3. Structural Characterization and Performance Testing

3.3.1. Structural Characterization

Infrared spectroscopy was conducted using Fourier transform infrared spectroscopy (FT-IR): samples and pure KBr were dried and ground to a particle size < 2 µm, then pressed into translucent sheets for measurement (scanning range 4000–400 cm−1).
Crystal structure of samples using X-ray diffraction (XRD): scanning range 5–90 °C; scanning rate 2 °C/min; step size 0.02; current 40 mV; voltage 40 kV.
The microstructure of the sample was analyzed using scanning electron microscopy (SEM): the sample was attached to a black conductive adhesive, and measured after vacuum gold spraying with an acceleration voltage of 1.0 kV.

3.3.2. Photocatalytic Performance Testing

Rhb solutions were prepared at mass concentrations of 5, 10, 15, 20, 25, 30, 35, 40, 45, 50 mgL−1. The absorbance (A) of each solution was measured using an UV–Vis spectrophotometer at 550 nm, and the equation of the standard curve of absorbance and mass concentration of Rhb (ρ1, mgL−1) was recorded as A = 0.0317ρ1 + 0.0012, R2 = 0.9996. Preparation of MO solutions at mass concentrations of 1, 2, 5, 10, 15, 20, 25, 30 mgL−1. The absorbance (B) of each solution was measured using an UV–Vis spectrophotometer at 465 nm, and the equation of the standard curve of absorbance and mass concentration of MO (ρ2, mgL−1) was recorded as B = 0.0489 ρ2 + 0.0048, R2 = 0.9997.
A xenon lamp was used to simulate solar light conditions, and Rhb (150 mL, 50 mg·L−1) and MO (100 mL, 20 mg·L−1) were added as the target pollutants, and 100 mg and 50 mg of g-C3N4 were added through their respective dark reactions for 60 min, and then appropriate amounts of 30% H2O2 solution and 0.1 mol·L−1 Na2S2O8 solution were added, respectively. The light source was turned on for the photocatalytic experiments, and samples were taken at different intervals, filtered through a 0.45 µm filter membrane, and then measured using an UV spectrophotometer. The photocatalytic and synergistic system for the degradation of anionic dyes are shown in Equations (11) and (12);
ln   ( C / C 0 ) = kt
φ / 100 = Δ C / C 0   ×   100 = [ ( C 0 C ) / C 0 ]   ×   100
wherein “C” is the mass concentration of TC at time t, in mg/L; “C0” is the mass concentration of initial TC, in mg/L; “φ” is the degradation rate (or adsorption rate), in %; “∆C” is the mass concentration of TC reduction after time t, in mg/L; “k” is the reaction rate constant, in min−1; and “t” is the reaction time, in min.

4. Conclusions

Compared with the pure g-C3N4, the degradation rates of MO and Rhb were increased by 2.5 and 3.5 times for the above two synergistic systems, respectively. Within 60 min of photocatalytic reaction, the degradation rates of MO and Rhb were 90.71% and 98.54% in the H2O2 synergistic system, respectively, and 92.61% and 99.56% in the Na2S2O8 synergistic system, respectively. It can be seen that the addition of an oxidant is beneficial to the effective separation and migration of photogenerated electron-hole pairs. The results of Na2S2O8 and H2O2 concentration on the degradation of pollutants showed that the photocatalytic degradation of the synergistic system showed a tendency to become better with the increase in the drop volume. Zeta potential and radical quenching experiments showed that ·OH and ·O2 were the main reactive radicals for the degradation of dyes in these synergistic systems. The experimental phenomena suggest that there are distinct pattern difference in the main reactive substances of anionic and cationic dyes in different synergistic systems, which need to be further investigated.

Author Contributions

Writing, review and editing, K.D.; writing-original draft, review and editing, W.Y.; supervision, G.C.; conceptualization, H.W. and X.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by [the Key Project of Provincial Natural Science Research in Anhui Universities] grant number [No. KJ2018 A0508], and [the Horizontal Project of University-Enterprise Cooperation] grant number [No. HYB20210234].

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.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Li, X.; Huang, G.H.; Chen, X.J.; Huang, J.; Li, M.G.; Yin, J.; Liang, Y.; Yao, Y.; Li, Y.P. A review on graphitic carbon nitride (g-C3N4) based hybrid membranes for water and wastewater treatment. Sci. Total Environ. 2021, 792, 148462. [Google Scholar] [CrossRef] [PubMed]
  2. Huang, L.Y.; Zhang, F.; Li, Y.; Ding, P.H.; Li, P.P.; Xu, H.; Li, H.M. Partial Oxidation of Sn2+ Induced Oxygen Vacancy Overspread on the Surface of SnO2−x/g-C3N4 Composites for Enhanced LED-Light-Driven Photoactivity. J. Inorg. Organomet. Polym. Mater. 2019, 29, 765–775. [Google Scholar] [CrossRef]
  3. Marimuthu, S.; Antonisamy, A.J.; Malayandi, S.; Rajendran, K.; Tsai, P.C.; Pugazhendhi, A.; Ponnusamy, V.K. Silver nanoparticles in dye effluent treatment: A review on synthesis, treatment methods, mechanisms, photocatalytic degradation, toxic effects and mitigation of toxicity. J. Photochem. Photobiol. B Biol. 2020, 205, 111823. [Google Scholar] [CrossRef] [PubMed]
  4. Sutar, S.; Patil, P.; Jadhav, J. Recent advances in biochar technology for textile dyes wastewater remediation: A review. Environ. Res. 2022, 209, 112841. [Google Scholar] [CrossRef] [PubMed]
  5. Ahmaruzzaman, M.; 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]
  6. Alias, N.H.; Jaafar, J.; Samitsu, S.; Ismail, A.F.; Mohamed, M.A.; Othman, M.H.D.; Rahman, M.A.; Othman, N.H.; Nor, N.A.M.; Yusof, N.; et al. Mechanistic insight of the formation of visible-light responsive nanosheet graphitic carbon nitride embedded polyacrylonitrile nanofibres for wastewater treatment. J. Water Process. Eng. 2020, 33, 101015. [Google Scholar] [CrossRef]
  7. Yan, X.; Wang, Y.Y.; Kang, B.B.; Li, Z.; Niu, Y.H. Preparation and Characterization of Tubelike g-C3N4/Ag3PO4 Heterojunction with Enhanced Visible-Light Photocatalytic Activity. Crystals 2021, 11, 1373. [Google Scholar] [CrossRef]
  8. Guo, R.T.; Wang, J.; Bi, Z.X.; Chen, X.; Hu, X.; Pan, W.G. Recent advances and perspectives of g–C3N4–based materials for photocatalytic dyes degradation. Chemosphere 2022, 295, 133834. [Google Scholar] [CrossRef]
  9. Liu, J.B.; Bao, Y.; Zheng, W.T. Analyses of some structural properties on a class of hierarchical scale-free networks. Fractals 2022, 30, 2250136. [Google Scholar] [CrossRef]
  10. Anwer, H.; Mahmood, A.; Lee, J.; Kim, K.H.; Park, J.W.; Yip, A.C.K. Photocatalysts for degradation of dyes in industrial effluents: Opportunities and challenges. Nano Res. 2019, 12, 955–972. [Google Scholar] [CrossRef]
  11. Shabir, M.; Yasin, M.; Hussain, M.; Shafiq, I.; Akhter, P.; Nizami, A.S.; Jeon, B.H.; Park, Y.K. A review on recent advances in the treatment of dye-polluted wastewater. J. Ind. Eng. Chem. 2022, 112, 1–19. [Google Scholar] [CrossRef]
  12. Liu, J.B.; Bao, Y.; Zheng, W.T.; Hayat, S. Network coherence analysis on a family of nested weighted n-polygon networks. Fractals 2021, 29, 2150260. [Google Scholar] [CrossRef]
  13. Singh, P.; Mohan, B.; Madaan, V.; Ranga, R.; Kumari, P.; Kumar, S.; Bhankar, V.; Kumar, P.; Kumar, K. Nanomaterials photocatalytic activities for waste water treatment: A review. Environ. Sci. Pollut. Res. 2022, 29, 69294–69326. [Google Scholar] [CrossRef]
  14. Qamar, M.A.; Shahid, S.; Javed, M.; Shariq, M.; Fadhali, M.M.; Madkhali, O.; Ali, S.K.; Syed, I.S.; Awaji, M.Y.; Khan, M.S.; et al. Accelerated Decoloration of Organic Dyes from Wastewater Using Ternary Metal/g-C3N4/ZnO Nanocomposites: An Investigation of Impact of g-C3N4 Concentration and Ni and Mn Doping. Catalysts 2022, 12, 1388. [Google Scholar] [CrossRef]
  15. Vignesh, S.; Eniya, P.; Srinivasan, M.; Sundar, J.K.; Li, H.T.; Jayavel, S.; Pandiaraman, M.; Manthrammel, M.A.; Shkir, M.; Palanivel, B. Fabrication of Ag/Ag2O incorporated graphitic carbon nitride based ZnO nanocomposite for enhanced Z-scheme photocatalytic performance of various organic pollutants and bacterial disinfection. J. Environ. Chem. Eng. 2021, 9, 105996. [Google Scholar] [CrossRef]
  16. Liu, J.B.; Zhao, J.; Min, J.; Cao, J. The Hosoya index of graphs formed by a fractal graph. Fractals 2019, 27, 1950135. [Google Scholar] [CrossRef]
  17. Song, J.H.; Zhao, K.; Yin, X.B.; Liu, Y.; Khan, I.; Liu, S.Y. Boosting the visible-light activity of ZrO2/g-C3N4 by controlling the crystal structure of ZrO2. J. Mater. Res. Technol. 2021, 36, 3086–3095. [Google Scholar]
  18. Lv, K.; Zhang, Q.Z. Nano-TiO2 photocatalytic technology and atmospheric pollution control. China Environ. Sci. 2018, 38, 852–861. [Google Scholar]
  19. Zeng, P.; Wang, Q.Y.; Zhang, H.J.; Wang, J.Q.; Chen, X.Y. Application of Photocatalysis techniques on treatment of wastewater with medium and low concentration of ammonium nitrogen. Sci. China Chem. 2019, 4, 78–82. [Google Scholar]
  20. Li, Y.J.; Cao, T.P.; Sun, D.W. A bismuth-rich Bi4O5Br2/TiO2 composites fibers photocatalyst enables dramatic CO2 reduction activity. Acta Mater. Compos. Sin. 2023, 41, 1–10. [Google Scholar]
  21. Meng, A.Y.; Zhang, L.Y.; Cheng, B.; Yu, J.G. Dual Cocatalysts in TiO2 Photocatalysis. Adv. Mater. 2019, 31, 1807660. [Google Scholar] [CrossRef] [PubMed]
  22. Yao, Y.F.; Yuan, J.Y.; Shen, M.; Du, B.; Xing, R. Synthesis and Photocatalytic Performance of ZnO Micro/Nano Materials Induced by Amphiphilic Calixarene. Chin. J. Inorg. Chem. 2022, 38, 261–273. [Google Scholar]
  23. Dong, J.Q.; Zhang, Y.; Hussain, M.I.; Zhou, W.J.; Chen, Y.Z.; Wang, L.N. g-C3N4: Properties, Pore Modifications, and Photocatalytic Applications. Nanomaterials 2022, 12, 121. [Google Scholar] [CrossRef] [PubMed]
  24. Du, J.; Wang, Z.; Li, Y.H.; Li, R.Q.; Li, X.Y.; Wang, K.Y. Establishing WO3/g-C3N4 Composite for “Memory” Photocatalytic Activity and Enhancement in Photocatalytic Degradation. Catal. Lett. 2019, 149, 1167–1173. [Google Scholar] [CrossRef]
  25. Tan, L.; Yu, C.F.; Wang, M.; Zhang, S.Y.; Sun, J.Y.; Dong, S.Y.; Sun, J.H. Synergistic effect of adsorption and photocatalysis of 3D g-C3N4-agar hybrid aerogels. Appl. Surf. Sci. 2019, 467, 286–292. [Google Scholar] [CrossRef]
  26. Palanivel, B.; Lallimathi, M.; Arjunkumar, B.; Shkir, M.; Alshahrani, T.; Al-Namshah, K.S.; Hamdy, M.S.; Shanavas, S.; Venkatachalam, M.; Ramalingam, G. rGO supported g-C3N4/CoFe2O4 heterojunction: Visible-light-active photocatalyst for effective utilization of H2O2 to organic pollutant degradation and OH radicals production. J. Environ. Chem. Eng. 2021, 9, 104698. [Google Scholar] [CrossRef]
  27. Guo, Y.G.; Sun, X.H.; Chen, Q.Q.; Liu, Y.J.; Lou, X.Y.; Zhang, L.; Zhang, X.J.; Li, Y.S.; Guan, J. Photocatalytic Applications of g-C3N4 Based on Bibliometric Analysis. Catalysts 2022, 12, 1017. [Google Scholar] [CrossRef]
  28. Ismael, M. A review on graphitic carbon nitride (g-C3N4) based nanocomposites: Synthesis, categories, and their application in photocatalysis. J. Alloys Compd. 2020, 846, 156446. [Google Scholar] [CrossRef]
  29. Saeed, M.; Muneer, M.; Haq, A.U.; Akram, N. Photocatalysis: An effective tool for photodegradation of dyes—A review. Environ. Sci. Pollut. Res. 2022, 29, 293–311. [Google Scholar] [CrossRef]
  30. Zhou, Y.M.; Elchalakani, M.; Liu, H.F.; Briseghella, B.; Sun, C.Z. Photocatalytic concrete for degrading organic dyes in water. Environ. Sci. Pollut. Res. 2022, 29, 39027–39040. [Google Scholar] [CrossRef]
  31. Pandey, S.K.; Mishra, P.K.; Tiwary, D. Enhanced photocatalytic performance of NiS/ZnO nanocomposite for the remediation of PNP and RhB dye. J. Environ. Chem. Eng. 2022, 10, 107459. [Google Scholar] [CrossRef]
  32. Ogata, F.; Yasuda, S.; Saenjum, C.; Nakamura, T.; Kawasaki, N. Interactions of Cationic and Anionic Dyes with Activated Carbons. e-J. Surf. Sci. Nanotechnol. 2020, 18, 269–274. [Google Scholar] [CrossRef]
  33. Asaithambi, S.; Sakthivel, P.; Karuppaiah, M.; Balaji, V.; Yuvakkumar, R.; Ravi, G. Visible light induced photocatalytic performance of Mn-SnO2@ZnO nanocomposite for high efficient cationic dye degradation. J. Mater. Sci. Mater. Electron. 2021, 32, 22168–22186. [Google Scholar] [CrossRef]
  34. Maruthupandy, M.; Muneeswaran, T.; Vennila, T.; Vaishali, C.V.; Anand, M.; Cho, W.S.; Quero, F. Photocatalytic efficiency of graphene/nickel oxide nanocomposites towards the degradation of anionic and cationic dye molecules under visible light. J. Photochem. Photobiol. A 2022, 427, 113819. [Google Scholar] [CrossRef]
  35. Ivanenko, I.; Hutsul, K.; Krymets, G. The precipitation synthesis of zinc (II) oxide for photocatalytic degradation of anionic and cationic dyes. Appl. Nanosci. 2022, 12, 755–759. [Google Scholar] [CrossRef]
  36. Gupta, T.; Chauhan, R.P. Enhanced photocatalytic degradation of cationic dye using Cu-doped ZnSe. Opt. Mater. 2023, 135, 113295. [Google Scholar] [CrossRef]
  37. Lu, E.J.; Tao, J.Q.; Yang, C.; Hou, Y.D.; Zhang, J.S.; Wang, X.C.; Fu, X.Z. Carbon-Encapsulated Pd/TiO2 for Photocatalytic H2 Evolution Integrated with Photodehydrogenative Coupling of Amines to Imines. Acta Phys. Chim. Sin. 2023, 39, 2211029. [Google Scholar] [CrossRef]
  38. Tan, F.X.; Chen, Y.K.; Wang, R.; Luo, C.W.; Wu, D.J.; Yang, L.B.; Zhang, Z. Performance and mechanism of g-C3N4/PDS for photocatalytic degradation of atrazine. China Water Wastewater 2023, 39, 91–98. [Google Scholar]
  39. Qu, Y.H.; Chen, Z.Y.; Duan, Y.K.; Liu, L.F. H2O2 assisted photocatalysis over Fe-MOF modified BiOBr for degradation of RhB. J. Chem. Technol. Biotechnol. 2022, 97, 2881–2888. [Google Scholar] [CrossRef]
  40. Su, H.L.; Christodoulatos, C.; Smolinski, B.; Arienti, P.; O’Connor, G.; Meng, X.G. Advanced Oxidation Process for DNAN Using UV/H2O2. Engineering 2019, 5, 849–854. [Google Scholar] [CrossRef]
  41. Liu, J.F.; Ran, Z.Z.; Cao, Q.Y.; Ji, S.F. Preparation of MIL-88B (Fex, Co1−x) catalysts and their application in one-step liquid-phase methanol oxidation to methyl formate using H2O2. Chin. J. Catal. 2021, 42, 2254–2264. [Google Scholar] [CrossRef]
  42. Shvalagin, V.V.; Kompanets, M.O.; Kutsenko, S.; Kuchmy, S.Y.; Skoryk, M.A. Photocatalytic Activity of g-C3N4 in the Partial Oxidation of Benzyl Alcohol Under Visible Light. Theor. Exp. Chem. 2020, 56, 111–116. [Google Scholar] [CrossRef]
  43. Yu, Y.T.; Huang, H.W. Coupled adsorption and photocatalysis of g-C3N4 based composites: Material synthesis, mechanism, and environmental applications. Chem. Eng. J. 2023, 453, 139755. [Google Scholar] [CrossRef]
  44. Liu, H.J.; Zhang, H.; Chen, M.; Zhang, C.N.; Du, C.W.; Peng, J.B.; Jiang, K. Peroxymonosulfate-assisted photocatalysis with g-C3N4/BiOCOOH nanocomposites for the synergistic removal of organic pollutants. J. Water Process. Eng. 2020, 38, 101580. [Google Scholar] [CrossRef]
  45. Lin, M.X.; Li, F.Y.; Wang, W.; Cheng, W.Y.; Zhou, C.L. Synthesis of biochar-supported P-doped g-C3N4 photocatalyst and its photocatalytic degradation mechanism to naphthalene. Huanjing Kexue Xuebao 2021, 41, 3200–3210. [Google Scholar]
  46. Ke, P.; Zeng, D.L.; Cui, J.W.; Li, X.; Chen, Y. Improvement in Structure and Visible Light Catalytic Performance of g-C3N4 Fabricated at a Higher Temperature. Catalysts 2022, 12, 247. [Google Scholar] [CrossRef]
  47. Qi, S.Y.; Liu, X.T.; Zhang, R.Y.; Zhang, Y.M.; Xu, H.Y. Preparation and photocatalytic properties of g-C3N4/BiOCl heterojunction. Inorg. Chem. Commun. 2021, 133, 108907. [Google Scholar] [CrossRef]
  48. Zhang, S.W.; Gao, H.H.; Liu, X.; Huang, Y.S.; Xu, X.J.; Alharbi, N.S.; Hayat, T.; Li, J.X. Hybrid 0D-2D Nanoheterostructures: In Situ Growth of Amorphous Silver Silicates Dots on g-C3N4 Nanosheets for Full-Spectrum Photocatalysis. ACS Appl. Mater. Interfaces 2020, 12, 3274. [Google Scholar] [CrossRef] [Green Version]
  49. Yang, H.H.; Wang, L.G.; Shuang Xu, S.; Cao, Y.; He, P.; Chen, J.Q.; Zheng, Z.; Li, H.Q. Green and selective hydrogenation of aromatic diamines over the nanosheet Ru/g-C3N4-H2 catalyst prepared by ultrasonic assisted impregnation-deposition method. Green Energy Environ. 2022, 77, 1361–1376. [Google Scholar] [CrossRef]
  50. Yao, G.Y.; Liu, Y.Q.; Liu, J.C.; Xu, Y. Facile Synthesis of Porous g-C3N4 with Enhanced Visible-Light Photoactivity. Molecules 2022, 27, 1754. [Google Scholar] [CrossRef]
  51. Liu, Y.F.; Ma, Z. TiOF2/g-C3N4 composite for visible-light driven photocatalysis. Colloids Surf. A Physicochem. Eng. Asp. 2021, 618, 126471. [Google Scholar] [CrossRef]
  52. Chen, Y.H.; Wang, B.K.; Hou, W.C. Graphitic carbon nitride embedded with graphene materials towards photocatalysis of bisphenol A: The role of graphene and mediation of superoxide and singlet oxygen. Chemosphere 2021, 278, 130334. [Google Scholar] [CrossRef]
  53. Ren, W.W.; Yang, J.K.; Zhang, J.X.; Li, W.; Sun, C.Y.; Zhao, H.L.; Wen, Y.T.; Sha, O.; Liang, B. Recent progress in SnO2/g-C3N4 heterojunction photocatalysts: Synthesis, modification, and application. J. Alloys Compd. 2022, 906, 164372. [Google Scholar] [CrossRef]
  54. Kumar, N.; Jung, U.J.; Jung, B.M.M.; Park, J.; Naushad, M. Zinc hydroxystannate/zinc-tin oxide heterojunctions for the UVC-assisted photocatalytic degradation of methyl orange and tetracycline. Environ. Pollut. 2023, 316, 120353. [Google Scholar] [CrossRef]
  55. Liu, Y.; Zhao, H.; Li, H.P.; Cai, T.F. Photocatalytic degradation of dyes by sulfur- and chlorine-co-doped g-C3N4 nanosheets. China Environ. Sci. 2021, 41, 4662–4669. [Google Scholar]
  56. Li, Z.Q.; Chen, X.W.; Wang, Y.; Cheng, T.; Huang, T.; Dong, P.Y.; Wang, W.Y.; Zhang, B.B.; Xi, X.G. Preparation of CeTiO4/g-C3N4 Composite with Efficient Photocatalytic Activity for Dye-Degradation. Chin. J. Inorg. Chem. 2022, 38, 53–62. [Google Scholar]
  57. Zhao, W.Y.; Yi, F.T.; Gan, H.H.; Zhang, H.N.; Qing, Y.X.; Jin, H.X.; Zhang, K.F. Synergistic Photocatalytic Dye Oxidation and Hexavalent Chromium Reduction by Chlorine Doped g-C3N4 Nanosheets. Mater. Rep. 2019, 33, 3377–3382. [Google Scholar]
  58. Cao, X.Y.; Tang, X.N.; Ma, H.; Chen, Y.Y. Study on TiO2@SiO2 Photocatalytic Degradation of Different Dyes and Its Mechanism. Rare Met. Mater. Eng. 2022, 51, 3001–3012. [Google Scholar]
  59. Yuan, H.; Shang, P.; Yang, J.; Huang, Q.; Song, L.; Jiang, X.F. Anion-Directed Self-Assembly of Calix [4] arene-Based Silver(I) Coordination Polymers and Photocatalytic Degradation of Organic Pollutants. Inorg. Chem. 2023, 62, 2652–2662. [Google Scholar] [CrossRef]
  60. Liu, F.; Nguyen, T.; Wang, Q.; Massuyeau, F.; Dan, Y.; Jiang, L. Construction of Z-scheme g-C3N4/Ag/P3HT heterojunction for enhanced visible-light photocatalytic degradation of tetracycline (TC) and methyl orange (MO). Appl. Surf. Sci. 2019, 496, 143653. [Google Scholar] [CrossRef]
  61. Song, Y.H.; Dai, W.L.; Xu, H.; Zhao, J.Z. Preparation and Photocatalytic Properties of g-C3N4/Bi12O17Cl2 Composites. Chin. J. Mater. Res. 2021, 35, 911–917. [Google Scholar]
  62. Ying, L.W.; Liu, Y.; Dong, W.B.; Yuan, H.X.; Ma, L.M. Effect of bromide ions on the UV/H2O2 oxidation of alanine in aqueous solutions. Chem. Eng. J. 2016, 285, 172–179. [Google Scholar] [CrossRef]
  63. Xu, H.; Ye, Q.; Wang, Q.G.; Zhou, P.; Huo, X.W.; Wang, Y.Q.; Huang, X.; Zhou, G.Y.; Zhang, J. Enhancement of organic contaminants degradation at low dosages of Fe (III) and H2O2 in g-C3N4 promoted Fe (III)/H2O2 system under visible light irradiation. Sep. Purif. Technol. 2020, 251, 117333. [Google Scholar] [CrossRef]
  64. Zhang, L.; Lin, Z.Y.; Zhang, W.J. Degradation of dyeing wastewater containing Rhodamine B by CuO-activated Persulfate with UV strengthening. Environ. Sci. Technol. 2020, 43, 82–89. [Google Scholar]
  65. Yang, L.H.; Ren, X.C.; Zhang, Y.J.; Chen, Z.Y. Heterogeneous activation of peroxymonosulfate by Cu+-decorated g-C3N4 under sunlight for degradation of organic pollutants. J. Environ. Chem. Eng. 2021, 9, 106596. [Google Scholar] [CrossRef]
  66. Zhang, Y.Q.; Li, L.F.; Zhou, F.; Yuan, X.Z.; Hu, S.Z.; Zhang, J. Zr-doped g-C3N4 photocatalytic degradation of organic pollutants. Fine Chem. 2022, 39, 2112–2121. [Google Scholar]
Figure 1. Infrared spectra and X-ray energy spectra of g-C3N4: (a) FT-IR; (b) XRD.
Figure 1. Infrared spectra and X-ray energy spectra of g-C3N4: (a) FT-IR; (b) XRD.
Molecules 28 02796 g001
Figure 2. SEM image of g-C3N4.
Figure 2. SEM image of g-C3N4.
Molecules 28 02796 g002
Figure 3. Rhb photocatalytic performance curves of different systems (photocatalyst dosage = 0.1 g, target pollutant = 50 mg/L): (a) photocatalytic activity of Rhb degradation photocatalyst under visible light irradiation; (b) the first-order kinetic curve of Rhb degradation photocatalyst under visible light irradiation.
Figure 3. Rhb photocatalytic performance curves of different systems (photocatalyst dosage = 0.1 g, target pollutant = 50 mg/L): (a) photocatalytic activity of Rhb degradation photocatalyst under visible light irradiation; (b) the first-order kinetic curve of Rhb degradation photocatalyst under visible light irradiation.
Molecules 28 02796 g003
Figure 4. MO photocatalytic performance curves of different systems (photocatalyst dosage = 0.05 g, target pollutant = 20 mg/L): (a) photocatalytic activity of MO degradation photocatalyst under visible light irradiation; (b) the first-order kinetic curve of MO degradation photocatalyst under visible light irradiation.
Figure 4. MO photocatalytic performance curves of different systems (photocatalyst dosage = 0.05 g, target pollutant = 20 mg/L): (a) photocatalytic activity of MO degradation photocatalyst under visible light irradiation; (b) the first-order kinetic curve of MO degradation photocatalyst under visible light irradiation.
Molecules 28 02796 g004
Figure 5. The absorption spectra of Rhb and MO under different catalytic conditions: (a) Rhb (g-C3N4); (b) Rhb (g-C3N4/H2O2); (c) Rhb (g-C3N4/Na2S2O8); (d) MO (g-C3N4); (e) MO (g-C3N4/H2O2); (f) MO (g-C3N4/Na2S2O8).
Figure 5. The absorption spectra of Rhb and MO under different catalytic conditions: (a) Rhb (g-C3N4); (b) Rhb (g-C3N4/H2O2); (c) Rhb (g-C3N4/Na2S2O8); (d) MO (g-C3N4); (e) MO (g-C3N4/H2O2); (f) MO (g-C3N4/Na2S2O8).
Molecules 28 02796 g005
Figure 6. Effect of different amounts of Na2S2O8 on the degradation rate of Rhb.
Figure 6. Effect of different amounts of Na2S2O8 on the degradation rate of Rhb.
Molecules 28 02796 g006
Figure 7. Effect of different amounts of H2O2 on the degradation rate of MO.
Figure 7. Effect of different amounts of H2O2 on the degradation rate of MO.
Molecules 28 02796 g007
Figure 8. Photocatalytic degradation of Rhb in the synergistic systems with different quenching agents: (a) Quenching of Rhb degradation in the H2O2 synergistic system; (b) Quenching of Rhb degradation in the Na2S2O8 synergistic system.
Figure 8. Photocatalytic degradation of Rhb in the synergistic systems with different quenching agents: (a) Quenching of Rhb degradation in the H2O2 synergistic system; (b) Quenching of Rhb degradation in the Na2S2O8 synergistic system.
Molecules 28 02796 g008
Figure 9. Photocatalytic degradation of MO in the synergistic systems with different quenching agents: (a) Quenching of MO degradation in the H2O2 synergistic system; (b) Quenching of MO degradation in the Na2S2O8 synergistic system.
Figure 9. Photocatalytic degradation of MO in the synergistic systems with different quenching agents: (a) Quenching of MO degradation in the H2O2 synergistic system; (b) Quenching of MO degradation in the Na2S2O8 synergistic system.
Molecules 28 02796 g009
Figure 10. The reusability of collaborative system:(a) g-C3N4/H2O2 (b) g-C3N4/Na2S2O8.
Figure 10. The reusability of collaborative system:(a) g-C3N4/H2O2 (b) g-C3N4/Na2S2O8.
Molecules 28 02796 g010
Figure 11. The FT-IR of synergetic system before and after degradation.
Figure 11. The FT-IR of synergetic system before and after degradation.
Molecules 28 02796 g011
Figure 12. Schematic diagram of the mechanism of synergistic photocatalysis.
Figure 12. Schematic diagram of the mechanism of synergistic photocatalysis.
Molecules 28 02796 g012
Table 1. Degradation rate of Rhb in synergistic systems.
Table 1. Degradation rate of Rhb in synergistic systems.
Time (min)Degradation Rate of
g-C3N4/H2O2 (100%)
Degradation Rate of
g-C3N4/Na2S2O8 (100%)
−6000
01.841.84
1043.5713.81
2064.6079.92
3080.9497.66
4087.1498.73
5093.3599.24
6098.5499.56
90100100
Table 2. Degradation rate of MO in synergistic systems.
Table 2. Degradation rate of MO in synergistic systems.
Time (min)Degradation Rate of
g-C3N4/H2O2 (100%)
Degradation Rate of
g-C3N4/Na2S2O8 (100%)
−6000
015.1815.18
1031.8744.96
2052.7565.73
3068.3386.51
6090.7192.61
9099.9098.70
15010099.80
210100100
Table 3. Comparison of the various parameters of dye degradation.
Table 3. Comparison of the various parameters of dye degradation.
CatalystQuality (mg)ContaminantsConcentration (mg/L)Volume (mL)Time
(h)
Degradation Rate (100%)Speed
(min−1)
ZHS/ZTO50MO10200268.45-
Sulfur/chlorine/g-C3N450Rhb1050--0.01683
CeTiO4/g-C3N4100Rhb101002.495.70.0202
Chlorine/g-C3N450Rhb10100--0.049
TiO2@SiO2100Rhb201001286-
TiO2@SiO2100MO201001238-
{[Ag2(mu-NO3) L1]} n10Rhb49.620685.20.00747
{[Ag2(mu-NO3) L1]} n10MO18.6820670.60.00354
This study
(g-C3N4)
100Rhb501501.51000.06907
0.10511
This study
(g-C3N4)
50MO201002.5
3.5
1000.07753
0.03978
Table 4. Zeta potential of g-C3N4 material.
Table 4. Zeta potential of g-C3N4 material.
MaterialsZeta Potential (mV)MaterialsZeta Potential (mV)
g-C3N4−14.4g-C3N4 (MO)−19.6
g-C3N4 (Rhb)−27g-C3N4 (MO, H2O2)−12.6
g-C3N4 (Rhb, H2O2)−20.3g-C3N4 (MO, Na2S2O8)−29.1
g-C3N4 (Rhb, Na2S2O8)−27.5
Table 5. Free radical capture control situation.
Table 5. Free radical capture control situation.
Degradation Rate (100%)Inhibition Rate (100%)
Noh+·O2·OHSO4·Noh+·O2·OHSO4·
Rhb (g-C3N4/H2O2)10099.7569.8535.4-00.2530.1564.6-
Rhb (g-C3N4/Na2S2O8)10010066.1295.1283.530033.884.8816.47
MO (g-C3N4/H2O2)10091.4120.0841.26-08.5979.9258.74-
MO (g-C3N4/Na2S2O8)10010051.3533.271000048.6566.730
Table 6. Main reactive radicals of different synergistic systems.
Table 6. Main reactive radicals of different synergistic systems.
TypeTarget PollutantSynergistic MultisystemMain Active Radical
cationic dyes (alkaline dye) RhbH2O2·OH
Na2S2O8·O2
anionic dyes (acid dyes)MOH2O2·O2
Na2S2O8·OH
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, W.; Ding, K.; Chen, G.; Wang, H.; Deng, X. Synergistic Multisystem Photocatalytic Degradation of Anionic and Cationic Dyes Using Graphitic Phase Carbon Nitride. Molecules 2023, 28, 2796. https://doi.org/10.3390/molecules28062796

AMA Style

Yang W, Ding K, Chen G, Wang H, Deng X. Synergistic Multisystem Photocatalytic Degradation of Anionic and Cationic Dyes Using Graphitic Phase Carbon Nitride. Molecules. 2023; 28(6):2796. https://doi.org/10.3390/molecules28062796

Chicago/Turabian Style

Yang, Wen, Kun Ding, Guangzhou Chen, Hua Wang, and Xinyue Deng. 2023. "Synergistic Multisystem Photocatalytic Degradation of Anionic and Cationic Dyes Using Graphitic Phase Carbon Nitride" Molecules 28, no. 6: 2796. https://doi.org/10.3390/molecules28062796

Article Metrics

Back to TopTop