Next Article in Journal
Enhancing Nickel-Iron Gas Diffusion Electrodes for Oxygen Evolution in Alkaline Water Electrolysis
Previous Article in Journal
Catalysts for Air Pollution Control: Present and Future
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Photocatalytic Degradation of Aqueous Atrazine over Graphene-Promoted g-C3N4 Nanosheets

Department of Soil Sciences & Agri-Food Engineering, Centre in Green Chemistry & Catalysis, Centr’Eau, Université Laval, Québec, QC G1V 0A6, Canada
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(9), 1265; https://doi.org/10.3390/catal13091265
Submission received: 25 July 2023 / Revised: 24 August 2023 / Accepted: 30 August 2023 / Published: 1 September 2023
(This article belongs to the Section Photocatalysis)

Abstract

:
Atrazine is a systemic herbicide widely used in weed control. In recent years, it has been largely detected in surface and groundwater in several locations all over the world. Photocatalysis is a green and sustainable technology with huge application prospects in pollution control and the degradation of organic water pollutants. In this work, photodegradation of aqueous atrazine was investigated over pristine graphitic carbon nitride (g-C3N4) synthesized via urea pyrolysis and graphene/g-C3N4 composite synthesized via the in situ growth method involving direct deposition of g-C3N4 nanosheets on the graphene surface. The obtained photocatalysts were characterized using transmission and scanning electron microscopy, Fourier-transformed infrared spectroscopy, UV-visible spectroscopy, photoluminescence spectroscopy, X-ray diffraction, and surface area measurements. It was demonstrated that the composite material exhibited remarkable photocatalytic properties for the efficient degradation of aqueous atrazine under visible light at ambient temperature. After 5 h of reaction, atrazine conversion reached 100% in the presence of graphene/g-C3N4 composite, while the pristine g-C3N4 allowed 40% conversion under the same conditions, thus demonstrating the positive effect of graphene on the photocatalytic activity of g-C3N4. Moreover, graphene/g-C3N4 was shown to keep its activity even when it was recycled five times, thus proving its stability and its potential to be used at the industrial scale.

Graphical Abstract

1. Introduction

With the continuous growth of the world’s population, the demand for crops has increased, and the use of chemical fertilizers and pesticides has seriously impacted global soil and water quality [1,2]. Pesticide residues in large quantities occurring in the environment have received much attention as new emerging contaminants. Herbicides, a class of pesticides, remain the most efficient and economical way to control weeds. With the development of herbicide-tolerant crops, the use of herbicides is increasing around the world, resulting in severe contamination of the environment [3]. Atrazine is a commonly used herbicide in several countries where it is among the most used and sold [4]. This herbicide is used on agricultural crops, in urban areas, in forestry and in railroad or energy transportation outlets. Many pesticides, mostly herbicides, including atrazine, have been detected in surface and groundwater [5]. Due to its high chemical stability, slow hydrolysis, low volatility and poor biodegradability, atrazine is often found and accumulates in soil ecosystems [6,7]. Despite atrazine’s low water solubility, there is much concern about the contamination of water resources [5,7]. Atrazine is recognized as an endocrine disruptor [6]. The metabolism of atrazine has been investigated in surrogate animals and humans [8]. Several investigations reported that atrazine’s half-life in the aquatic environment and groundwater resources ranges from 41 to 237 days and 15 months to 20 years, respectively [9]. The maximum contaminant level for atrazine in drinking water established by the United States Environmental Protection Agency (US EPA) and the World Health Organization (WHO) is 3.0 and 2.0 μg·L−1, respectively [10]. It is, therefore, essential to find effective solutions for the degradation of such water pollutants.
Photocatalysis, an advanced oxidation process (AOP), is considered as a cost-effective means to degrade toxic and harmful pollutants. In recent years, graphitic carbon nitride (g-C3N4), a metal-free visible-light-activated semiconductor, has attracted significant attention as a promising material for various environmental remediation applications and pollutant degradation [11]. Due to its unique electronic band structure, g-C3N4, a low-cost and non-metal photocatalyst, shows great application prospects in the decomposition of water into hydrogen and oxygen, photocatalytic degradation of organic pollutants and carbon dioxide reduction. However, g-C3N4 exhibits some shortfalls, such as small specific surface area, low light quantum yield, and high recombination of electron–hole pairs, thus reducing its application in the field of photocatalysis [12]. To improve the photocatalytic activity of g-C3N4, many strategies have been developed. They mainly included developing various nanostructures (such as nanosheets, nanoparticles, nanorods and nanospheres) [13], as well as morphological [14,15,16,17], molecular [18,19,20,21,22], and electronic [23,24,25] modifications. Also, surface engineering strategies such as cocatalyst loading and heterogeneous structure construction have been reported [26,27]. Graphene is a two-dimensional structure of sp2-hybridized carbon atoms arranged hexagonally in a honeycomb framework with a high strength-to-weight ratio, unique and extraordinary electrical properties, high thermal conductivity, and high surface area [28,29,30,31,32]. The electronic–band structure of graphene combines semiconducting and metallic characteristics [28]. This property stems from graphene’s unit cell, comprised of two carbon atoms, which makes the π and π* bands indistinguishable. Moreover, the close carbon–carbon distance in graphene leads to a pronounced interatomic overlap [33,34]. Thus, for application purposes, graphene is not used as pristine, but it must be employed as reduced graphene oxide, activated graphene, doped graphene, graphene/metal oxide composites or graphene/polymer composites [30]. Since these carbon materials exhibit high electron mobility, the photocatalytic activity of g-C3N4 can be significantly enhanced by coupling with these carbon materials. However, it has been shown that it is difficult to directly use pristine graphene as a material due to its long and difficult synthesis as well as the formation of agglomerations in solution due to van der Waals interactions [35,36,37]. An alternative approach is the recourse to graphene oxide (GO) compound with a similar structure of graphene (G) and presenting hydroxyl (–OH), alkoxyl (C–O–C), carbonyl (C=O), carboxylic acid (–COOH) and other oxygen-based surface functional groups [38]. Graphene oxide has many advantages and disadvantages. Among the advantages, GO is easily synthesized, and its oxygen groups allow not only a very high wettability but also the possibility of functionalization of its surface. Also, it is feasible to synthesize reduced graphene oxide (rGO) or graphene (G) using different methods, with the aim of minimizing the number of oxygen groups and obtaining properties closer to those of pristine graphene [39]. However, these electrical, thermal, and mechanical properties sometimes limit its application. Therefore, when synthesizing GO, the excessive addition of functional groups in its surface impairs its electrical conductivity, making GO electrically resistive (1.64 × 104 Ω·m). In addition, GO has a low thermal conductivity of 0.5 to 1 W/m·K [40]. Given these modest properties, several researchers have explored various techniques to improve the performance of GO. For instance, the reduction of GO to graphene (G) or reduced graphene oxide (rGO) seems to be an effective alternative to obtain excellent material properties. Several published works reported a very clear overview of the properties of graphene compared to those of GO. Graphene has a high electron mobility (25 m2/V·s) as well as a high electrical conductivity (6500 S/m) [41,42]. Indeed, the reduction of GO via a pyrolysis method at very high temperatures is based on the decomposition of oxygen groups into CO and CO2 gases [39]. These gases effectively exfoliate the graphene oxide into nanosheets.
The combination of graphene with graphitic carbon nitride into composites with improved photocatalytic efficiency was previously reported for the photodegradation of dyes such as methylene blue, methyl-orange, and rhodamine [43]. However, graphene/graphitic carbon nitride composites were hardly reported for the photocatalytic degradation of pesticides in general and atrazine in particular [44]. Moreover, few g-C3N4-based photocatalysts were reported for atrazine degradation, such as AgCl@pg-C3N4 nanocomposites [45], ternary MoS2-loaded ZnO-g-C3N4 nanocomposites [46], Z-scheme CdS/RGO/g-C3N4 hybrid photocatalysts [47], activated carbon/g-C3N4 composites with peroxymonosulfate activation [48], oxygen-doped g-C3N4 [49] and p-p stacked step-scheme PDI/g-C3N4/TiO2@Ti3C2 [18].
The aim of the present work is the investigation of the visible light-driven photocatalytic degradation of aqueous atrazine in the presence of g-C3N4 and the composite graphene/g-C3N4 for the first time in the literature. The specific objectives are (i) the synthesis and characterization of the photocatalytic materials and (ii) the investigation of the effect of operating conditions on the reaction advancement.

2. Results and Discussion

2.1. Photocatalysts Characterization

The XRD patterns of the synthesized photocatalysts, i.e., pure g-C3N4 and graphene/g-C3N4 composites, are shown in Figure 1. There are two peaks of reflections for both materials. The peaks are most pronounced and centered around 13.1° and 27.4°, indicating very clearly that the graphitic-type crystal structure of g-C3N4 was formed [50,51]. The low-intensity peak appearing at 2θ = 13.1° can be indexed to the (100) plane and corresponds to a structural rearrangement pattern between the tri-s-triazine (heptazine) continuous planes with an interlayer distance of 0.676 nm [52,53]. Moreover, the strong peak at 2θ = 27.4° is attributed to a (002) plane, reflecting the periodic stacking of the conjugated aromatic structure with an interlayer distance of 0.33 nm [54,55,56]. Also, another low-intensity diffraction peak is observed at 2θ = 21.8° with a (002) plane, probably due to the low amount of graphene oxide remaining after the reduction of GO to graphene. Interestingly, Sengupta et al. [57] also reported the occurrence of two characteristic peaks corresponding to (002) and (100) planes in their investigation of the effect of temperature on the reduction of graphene oxide to graphene. These findings are comparable with the XRD data reported in the present work; therefore, it can be concluded that the photocatalysts under study were well synthesized and had a high structural crystallinity corresponding to pure g-C3N4 and graphene/g-C3N4. The presence of two diffraction peaks always existed in graphene/g-C3N4 with 2θ = 13.1° and 2θ = 27.4° implying a strong interaction between graphene and g-C3N4 [58].
Figure 2 depicts the SEM and TEM images of g-C3N4 and graphene/g-C3N4 photocatalysts. The SEM analysis of the g-C3N4 material revealed a typical flaky and ultrafine nanosheet structure (see Figure 2a). These nanosheets are curled to reduce the surface tension. This particular property is identical to the construction of the exfoliated carbon scrolls during graphite peeling [59,60]. The graphene/g-C3N4 composite exhibited a sandwich structure, which could explain the very close contact between the g-C3N4 and graphene (Figure 2b) [50,61]. Also, in the enlarged image of the targeted location (Figure 2b inset), it is clearly seen that the g-C3N4 lamellas are dispersed on the graphene surface, which may not only minimize the aggregation of g-C3N4 but may also increase the active sites of graphene/g-C3N4 for the photocatalytic reaction [58]. As for the TEM analysis, g-C3N4 exhibited a typical small, thin and flat irregular two-dimensional shape (see Figure 2c), while graphene/g-C3N4 showed no free graphene sheets (see Figure 2d) apart from the g-C3N4 sheets. This may suggest that the in situ growth method successfully integrated g-C3N4 with graphene.
Figure 3 depicts the FTIR spectra belonging to g-C3N4 and graphene/g-C3N4 materials. In the case of g-C3N4 material, a broad band around 3600 cm−1 is observed and assigned to N–H stretching vibrations of non-condensed amino groups in the structure of g-C3N4 [54,62]. Moreover, the broad, strong peak around 3350 cm−1 was attributed to hydroxyl groups (–OH) and physically adsorbed water [58]. The strong peaks at 1235, 1325, 1403, 1572 and 1637 cm−1 correspond to the typical stretching vibrations of (C=N) and (C–N) [52]. In addition, the band around 812 cm−1 corresponds to the bending modes of the tri-s-triazine backbone units [52]. For graphene/g-C3N4, the peak at 3350 cm−1 (–OH) flattened slightly, thus signifying a partial removal of hydroxyl groups (–OH) [59]. Since the synthesis of graphene/g-C3N4 involves an in situ thermal reduction of the graphene oxide, the disappearance of characteristic absorption peaks belonging to graphene oxide may indicate the effective GO reduction, as previously reported [57]. The absorption peak at 1734 cm−1 corresponding to the stretching vibrations of carbonyl groups (–C=O–) of GO was not observed. However, a weak peak at 1637 cm−1 attributed to an epoxy group (–CO–) known by its high binding energy occurred, thus demonstrating that it is difficult to remove all the epoxy in graphene oxide following the thermal reduction [58]. The weak intensity of the band at 1403 cm−1 was attributed to a partial removal of the carboxylic groups (–COOH) [57]. Furthermore, the peak at 1325 cm−1 is responsible for the C=C backbone vibration in graphene. In summary, it can be concluded that the characteristic peaks of g-C3N4 did not change after graphene deposition.
Textural properties of the g-C3N4 and graphene/g-C3N4 materials were investigated using N2 adsorption–desorption volumetric analysis. As illustrated in Figure 4, the obtained isotherms were attributed to the type IV with H3 hysteresis loop according to the IUPAC classification, thus indicating the occurrence of a mesoporous structure. The BET-specific surface areas of g-C3N4 and graphene/g-C3N4 were 54.4 and 74.8 m2/g, respectively. The occurrence of mesopores was corroborated by the pore diameters evaluated to be 3.4 and 4.9 nm for g-C3N4 and graphene/g-C3N4, respectively. The pore volumes reached 0.37 and 0.68 cm3/g for g-C3N4 and graphene/g-C3N4, respectively.
PL spectroscopy is a well-known technique to study both the interface charge carrier transfers and the electron–hole pair recombination. The PL emission spectra were thus employed to investigate the combination and separation of the photogenerated charges. Generally, the stronger PL intensity indicated the faster recombination of photogenerated charge carriers [63]. Figure 5a shows the PL spectra of g-C3N4 and graphene/g-C3N4. As seen, the PL intensity of graphene/g-C3N4 was lower compared with pure g-C3N4, implying that the recombination rate of electron–hole pairs is substantially slowed, thus improving charge separation. Meanwhile, the emission peaks maximum red-shifted from 494 nm to 528 nm, which could promote the π-electron delocalization of the graphene/g-C3N4 network. This is associated with the decreased band gap energy [63]. Time-resolved fluorescence spectroscopy of g-C3N4 and graphene/g-C3N4 was also performed to study the charge separation dynamics (See Figure 5b). The lifetime of the charge carriers is 4.18 ps for g-C3N4 material. Interestingly, this lifetime increased to 39.89 ps for graphene/g-C3N4, which is almost 10 times that of the charge carriers of g-C3N4, indicating a more efficient separation of electrons and holes in graphene/g-C3N4 compared to pure g-C3N4. Therefore, these photogenerated carriers of graphene/g-C3N4 were more stable and able to initiate redox reactions prior to recombination [64]. This behavior impacts the photocatalytic performance, which depends mainly on the number of active sites, the extent of visible light absorption and the efficiency of charge carrier generation–separation and transfer. As reported by Dong et al. [65], the improvement in the photocatalytic activity brought by the heterojunctions such as graphene/g-C3N4 is attributed to the improvement of the separation of electrons and photogenerated holes. The large number of active sites and direct transport or transfer pathways for fast charge carriers within nanostructures could significantly improve charge carrier separation.
The registered UV-Vis spectra are depicted in Figure 6a. These UV-Vis spectra were performed to characterize the visible light absorption capacity and optical band gap energy of g-C3N4 and graphene/g-C3N4, respectively. From Figure 6a, it is found that g-C3N4 effectively absorbs in the visible light region, and the absorption edge was at about 494 nm. After the incorporation of graphene into g-C3N4, the graphene/g-C3N4 photocatalyst showed a significantly enhanced absorption in the visible light region at about 528 nm. As shown in (Figure 6b), Tauc plots based on the following equation were used to evaluate the band gap energies:
α h ν = A h ν E g n / 2
where α represents the absorption coefficient, h is Planck’s constant, ν is the frequency of light, A is a constant that it is independent of energy, Eg is the band gap energy of a photocatalyst, and n = 4 for direct band gap of g-C3N4 [66].
Given that the value of n of graphitic carbon nitride is 4, the Eg of photocatalysts could be determined by the tangent lines versus (hν) [66]. Therefore, the obtained results indicated that the band gap energies of g-C3N4 and graphene/g-C3N4 were 2.51 and 2.35 eV, respectively. The band gap of g-C3N4 compares relatively well with that found in the literature [12]. These findings confirm that the combination of g-C3N4 semiconductor with graphene can decrease the band gap energy of the composite material and significantly increase the absorption in the visible light region, which could promote higher photocatalytic activity.

2.2. Photocatalytic Reaction

To assess the adsorption and photocatalytic performances, the removal of aqueous atrazine was carried out under dark and visible light irradiation in the presence of the photocatalysts. The obtained results indicated that the equilibrium was reached within ca. 20 min. Performances of g-C3N4 and graphene/g-C3N4 were investigated under similar conditions, as presented in Figure 7. Noticeably, the adsorption under dark conditions allowed the removal of ca. 25 and 50% of atrazine in the presence of g-C3N4 and graphene/g-C3N4, respectively. The subsequent photocatalytic degradation performed under visible light allowed to reach ca. 50 and 100% removal of atrazine within 5 h of reaction over g-C3N4 and graphene/g-C3N4, respectively. Clearly, the composite containing graphene largely outperformed the pristine g-C3N4. This improved photocatalytic activity under visible light may be attributed to the important role of graphene in the enhanced photoinduced electron conduction through the graphene sheets. Additionally, its two-dimensional, planar π-conjugation structure gives graphene excellent electron conductivity, allowing for rapid transport that results in efficient charge separation [67].
The effect of photocatalyst loading on the photodegradation of atrazine is presented in Figure 8a. The results showed that the photocatalytic conversion of atrazine ranged from 54% to 100% within 5 h of the reaction when the photocatalyst loading was increased from 0.05 to 0.4 g/L. The effect is due to the increase in the amount of photocatalyst, which increases the number of active sites of graphene/g-C3N4 photocatalyst. Interestingly, it was found that the loading of 0.4 g/L was optimum, leading to the best performances (see Figure 8b). Indeed, the photocatalytic degradation of atrazine decreased when the photocatalyst loadings were set beyond the optimum loading of 0.4 g/L, i.e., 0.6, 0.8 and 1 g/L. This effect may be attributed to the fact that the degradation reaction was hampered by the shielding effect occurring in the reactor when high photocatalyst loadings are used. Indeed, under these conditions, agglomeration of the photocatalyst particles may occur, thus generating an obstacle preventing light from reaching the surface of the photocatalyst. Therefore, an optimal photocatalyst loading is likely able to generate more hydroxyl radicals and superoxide reactive species, leading to better photocatalytic degradation of atrazine than high loadings [68,69]. Sharma et al. also noted that the excessive photocatalyst loadings during the reaction led to a decrease in photocatalytic activity [70].
pH is among the most important factors affecting the efficiency of many chemical and biological processes [71]. It also plays an important role in the production of hydroxyl radicals (OH•), which have been shown to oxidize many different recalcitrant organic pollutants into mineral end products [72]. The photocatalytic reaction medium pH referring to a specific concentration of protons (H+) or hydroxyl (OH) ions in aqueous solutions can have an impact on the protonation or deprotonation of pollutants as well as photocatalysts surfaces, thus influencing the ability of the photocatalyst to adsorb the reactants, its activity, and the course of the reaction [73]. Therefore, the effect of pH on the degradation of atrazine in the presence of graphene/g-C3N4 composite was investigated. The original atrazine solution had a pH of 5.33, and the addition of hydrochloric acid and sodium hydroxide in solution regulated the other initial pH values (3.3, 2.5, 7.0, 8.6, 9.0 and 11.7). As shown in Figure 9, the conversion of atrazine decreased with increasing initial pH except for pH = 5.3, where there was no noticeable difference in the degradation of atrazine under acidic conditions. Rapid degradation occurred at low initial acidic pH at (pH = 3.3) and high at (pH = 5.3), where pH was the optimum, and where 85.2% and 100% of atrazine, respectively, was removed within 5 h of exposure time. Therefore, it can be concluded that acidic conditions are advantageous for the photocatalytic degradation of atrazine. Indeed, more H+ will bind to the surface of the graphene/g-C3N4 photocatalyst under acidic conditions, positively charging the surface. The single chloride atom of atrazine has an electron-withdrawing group and high electronegativity, allowing stronger adsorption by the photocatalyst. Xue et al. reported a BiOBr/UiO-66 composite which was very effective for the degradation of atrazine under visible light; they reported that this composite has the same reaction impact in view of the degradation with pH effect as the graphene/g-C3N4 composite. Their results show that the positively charged composite could enhance the electron transport on the surface of the photocatalyst and thus react with oxygen to produce a superoxide radical (O2). In addition to this advantage, the recombination of charge carriers (e/h+) could be decreased, and thus, the electron reacts with the with O2 from the solution, obtaining (O2) and efficiently oxidizing atrazine in the presence of graphene/g-C3N4 [74].
The heterogeneous photocatalysis allows the conversion of photonic energy (solar or artificial radiation) into chemical energy [75]. The photocatalyst (graphene/g-C3N4) absorbs visible light photon energy greater than or equal to its band gap energy. The electrons are excited to move from the valence band (BV) to the conduction band (BC), leading to the formation of electron–hole pairs (see Figure 10). The free charge carriers of the photocatalyst are positively charged holes (h+) in the valence band (BV) and electrons (e) in the conduction band (BC). These charge carriers, the electron–hole pairs in the BC and BV, respectively reduce and oxidize the compounds adsorbed on the surface of the photocatalyst. However, the photogenerated electron–hole pairs have a very short lifetime of the order of 10−9 s, which means that the probability of the chemical reaction time of the photocatalyst with atrazine is very low. The rapid recombination strongly affects the photocatalytic activity. The mechanism of heterogeneous photocatalysis of a g-C3N4 deposition on graphene can be explained as follows: (1) increased absorption of visible light, (2) separation and transfer of photoinduced charge carriers to the surface of the photocatalyst, and (3) reaction between the charge carriers (e/h+) and atrazine molecules. A schematic diagram of the mechanism of photodegradation of atrazine is shown in proposed in Figure 10). According to Du et al. [67], electrons can be directly excited from the valence band of graphene to the conduction band of g-C3N4. Indeed, graphene is a conductor that allows electrons to move freely. The energy of g-C3N4 in the conduction band is more negative than that of graphene. Therefore, photoinduced electrons can be transferred from the conduction band of g-C3N4 to graphene. The electrons of g-C3N4 can reduce the oxygen O2 present in the solution into active superoxide O2 since the redox potential NHE (O2/O2) is −0.33 eV). These electron–hole interlayer transfers between g-C3N4 and graphene extend the lifetime of the charge carriers. The valence band energy of g-C3N4 is more positive than the valence band energy of graphene, but the valence band energy of g-C3N4 (+1.57 eV) is less positive than the potential of (•OH/H2O = +1.99 eV), which does not allow the EVB holes of g-C3N4 to oxidize H2O and produce OH• hydroxyls. The adsorbed atrazine on the surface of the photocatalyst decomposes by transferring electrons (e) to the positive holes (h+) on the valence band of graphene/g-C3N4. These charge carriers trigger oxidation/reduction reactions that degrade the adsorbed contaminants (atrazine) and transform them into harmless products such as CO2, H2O, and other by-products such as inorganic acids. The equations of the photocatalytic reaction of atrazine over graphene/g-C3N4 are proposed together with the illustration of this mechanism (Figure 10), in agreement with the findings reported in the literature [50].
G r a p h e n e + g C 3 N 4 + h ν ( e ) + h + Generation   of   electron hole   pairs
e + O 2 O 2 Oxygen   reduction   to   superoxide
g C 3 N 4 e + O 2 g C 3 N 4 e + O 2 Reduction   of   oxygen   to   superoxide
A t r a z i n e + O 2 + C O 2 + H 2 O + o t h e r   e n d p r o d u c t s atrazine   degrades   photocatalytically   with   superoxide   radicals .
Stability and reusability are crucial for a heterogeneous photocatalyst. To assess the stability of the graphene/g-C3N4 photocatalyst in the photocatalytic degradation of atrazine, five reaction cycles were carried out under the optimum conditions of the above-mentioned reactions, i.e., photocatalyst loading of 0.4 g/L in contact with 10 mg/L atrazine solution, reaction duration 5 h and pH = 5.3, using the same photocatalyst batch. After each run, the photocatalyst was collected and rinsed several times with distilled water and dried for 24 h before the beginning of the next cycle. As depicted in Figure 11, it was found that the graphene/g-C3N4 composite, as designed, reveals stability and reusability after five cycles of recycling, as there was no significant loss of photocatalytic activity for atrazine degradation. It can be concluded that the graphene/g-C3N4 photocatalyst is of good quality as it maintains excellent stability and photocatalytic activity. Additionally, the conditions of its reuse require only a simple rinsing with pure water followed by drying, thus facilitating its long-lasting and efficient reuse.
Table 1 presents recent investigations reported in the literature regarding the photocatalytic degradation of aqueous atrazine using graphitic carbon nitride-based photocatalysts. Compared to the reported findings, the graphene/g-C3N4 photocatalyst performed very well, as complete degradation of atrazine was reached within a reasonable reaction time. Moreover, even though the investigations reported in Table 1 were promising, the procedures used for the modification of the graphitic carbon nitride were either complicated and time-consuming or involved the use of toxic metals such as molybdenum and cadmium, which may leach in the reaction medium, thus leading to serious environmental threats. Additionally, it is worth mentioning that the solubility of atrazine in water is around 30 mg/L at 20 °C [76]; therefore, the initial concentration of 100 mg/L reported by Mohamed (2015) is certainly inaccurate. Moreover, in the case of the pyromellitic diimide-modified g-C3N4/g-C3N4 homojunction, peroxymonosulfate was used to activate the photocatalyst (see reference [18]).

3. Experimental Methods

3.1. Materials

The chemicals used in this study were of a very high analytical quality and without the need for further purification. Atrazine (C8H14ClN5, 99%) and terbuthylazine (C9H16ClN5, 99%) were supplied by Sigma-Aldrich (Oakville, ON, Canada). Urea (CH4N2O, 99%) was purchased from BDH Chemicals (Toronto, ON, Canada). The solvent acetonitrile (C2H3N, 99%), anhydrous magnesium sulfate (MgSO4, 98%), hydrogen chloride (HCl, 1N) and sodium hydroxide (NaOH, 1N) were acquired from Fisher ScientificChemicals (Ottawa, ON, Canada). For GO (graphene oxide) synthesis, natural graphite flakes (325 mesh, 99.8%) from Alfa Aesar (Ottawa, ON, Canada); potassium persulfate (K2S2O8 ≥ 99%) from Anachemia (Montréal, QC, Canada); phosphorus pentoxide (P2O5 ≥ 99%) from JT Baker (Mississauga, ON, Canada); sulfuric acid (H2SO4, 98%) and potassium permanganate (KMnO4 ≥ 99%) from Fisher Scientific Chemicals (Ottawa, ON, Canada); hydrogen peroxide solution (H2O2, 30% by weight in water) from EMD Millipore (MilliporeSigma, Oakville, ON, Canada); hydrochloric acid (HCl, 36%) from Fisher ScientificChemicals (Ottawa, ON, Canada) and ethylene glycol (≥99%) from BDH (West Chester, PA, USA) were used. Millipore water (18 MΩ·cm) was used in the present investigation.

3.2. Synthesis of Pure g-C3N4 and Graphene/g-C3N4 Composite

GO was first prepared as reported by Yamaguchi et al. [78]. The composite graphene/g-C3N4 was synthesized as reported by Zhu et al. [79]. In a typical synthesis, 4 g of urea were dissolved in 50 mL of deionized water and put under magnetic stirring for 1 h. Then, 2 mg of GO was suspended in 50 mL of deionized water and ultrasonicated for 4 h so that a complete dispersion of the GO sheets was obtained. The urea solution and the GO suspension were then mixed. The mixture was ultrasonicated for 3 h and magnetically agitated for 3 h, and then dried in the oven at 80 °C for 24 h. The obtained solid was then calcinated at 550 °C in a covered crucible for 2 h at a heating rate of 10 °C·min−1. The final product of a gray color was denoted graphene/g-C3N4. The synthesis of pristine g-C3N4 was performed using urea as a precursor. Typically, 5 g of urea was dissolved in 100 mL of deionized water under ultrasonic and magnetic stirring for 1.5 and 1 h, respectively. The solution was dried at 60 °C for 12 h, then calcinated at 550 °C in a covered crucible for 2 h at a heating rate of 10 °C·min−1. The obtained yellow powder was designated g-C3N4.

3.3. Photocatalysts Characterization

The crystal structure of the as-prepared samples was analyzed using X-ray diffractometer Rigaku MiniFlex 6G (The Woodlands, TX, USA) operating with copper radiation (Kα = 1.5406 Å) at a power of 600 W (40 kV, 15 mA). Powder diffraction patterns were obtained between 5° and 60° with a scan speed of 5°/min. The morphology and size were examined via scanning electron microscopy (SEM) using a JEOL 6360LV microscope (Tokyo, Japan) with an acceleration voltage ranging from 5 to 25 kV, with the Secondary Electron Imaging signal (SEI). For the Transmission Electron Microscopy (TEM), a JEOL 1230 (Tokyo, Japan) equipped with a Gatan 1000XP bottom-mounted camera was used with an acceleration voltage of 80 kV. A HP UV-8453 UV-visible spectrophotometer using MgO as a background reference was used to record the UV–visible diffuse spectra of the synthesized samples. The photoluminescence (PL) spectra were obtained using a fluorescence spectrometer Horiba Quanta-Master 8000 (Edison, NJ, USA). Horiba scientific software was used to determine the rate of electron transfer. A decay mathematical approach was used, where τ represents the measured lifetime value. Fourier-transform infrared spectra (FTIR) were recorded between 300 and 4000 cm−1 on a Varian 1000 spectrometer Scimitar series (Foster City, CA, USA) with a resolution of 2 cm−1 using KBr pellets at room temperature. The textural properties of the samples were studied using nitrogen adsorption/desorption analyses at 77 K using a volumetric adsorption analyzer (model Autosorb-1, Quantachrome Instruments, Boyton Beach, FL, USA). The specific surface area was evaluated from the linear part of the Brunauer–Emmett–Teller (BET) plot (P/P0 = 0.05–0.30). Pore size distribution curves were determined from the desorption branch using the Barrett–Joyner–Hallenda (BJH) method. The total pore volume was estimated from the amount adsorbed at a relative pressure P/P0 = 0.99 single point.

3.4. Photocatalytic Activity Tests

The aqueous atrazine (10 mg/L) was photodegraded at ambient temperature in a 48 W photocatalytic reactor operating under visible light at a wavelength λ > 420 nm. The tests were operated batchwise under the following ranges of conditions: photocatalyst loading 0.05–1 g/L, pH 3.3–11.7 and time 0–300 min. The reacting volume was 5 mL for all the tests performed under agitation (750 rpm). To avoid any disturbance due to multiple liquid sampling on kinetic measurements, a separate experiment was used for each measurement. Furthermore, tests were performed in the dark to assess the adsorption occurring on the photocatalysts. At pre-set reaction times, aliquots of the aqueous solution were withdrawn, filtered with a microsyringe (0.2 μm), solvent extracted and analyzed for atrazine content using a HP 6890 Series gas chromatograph equipped with a flame ionization detector (Thermo Fisher Scientific, Markham, ON, Canada). The column used was an HP-5 (30 m × 0.25 mm ID × 0.25 μm film thickness). The injector and detector temperatures were set to 250 °C and 1 μL sample was injected with a split ratio of 50:1. The oven temperature profile was: initial temperature 40 °C for 2.5 min, ramp at 25 °C/min to 200 °C, hold for 5 min; ramp at 25 °C/min to 250 °C, hold for 1 min.
In addition, recycling experiments were performed for five consecutive cycles to test the stability and reusability of the photocatalysts. After each cycle, the photocatalyst was separated via centrifugation and gently washed with pure water several times to remove any residual atrazine impurities and then dried at 80 °C for the next test.
All the tests reported in the present investigation were performed in duplicate and analyzed twice.

4. Conclusions

Photodegradation of atrazine was achieved using urea-based graphitic carbon nitride (g-C3N4) and graphene from graphene oxide pyrolysis. The photocatalytic activity of graphene/g-C3N4 was found to be better than that obtained via bare g-C3N4 towards the degradation of atrazine. This was attributed to the fact that graphene increases the lifetime of electron–hole charge carriers (e/h+) and improves the creation of adsorption sites in the graphene/g-C3N4 network, thus favoring the reduction of the energy gap by increasing the population of electrons present in the solution. It was also noticed that this reduces the recombination of electron–hole pairs and improves the charge separation. The overall results showed a complete degradation of 100% of atrazine within 300 min at room temperature and under visible light irradiation, even after five cycles of photocatalyst recycling. The effects of pH conditions and photocatalyst loading on the degradation extent of atrazine using optimally synthesized graphene/g-C3N4 were investigated. Interestingly, the normal (uncontrolled) pH value of the atrazine solution and a moderate catalyst loading of 0.3 g/L led to the best degradation performances.

Author Contributions

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

Funding

This research was funded by the Natural Sciences and Engineering Research Council of Canada (NSERC) (Grant No: 2014-05355).

Data Availability Statement

All the data generated or analyzed within the present investigation are included in this manuscript.

Acknowledgments

Funding provided by the Natural Sciences and Engineering Research Council of Canada (NSERC) and the Canada Foundation for Innovation (CFI) is gratefully acknowledged. The authors would like to thank T.O. Do for the use of the UV-Vis spectrometer.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, L.; Chen, H.; Li, Y.; Wang, S.; Su, J.; Liu, X.; Chen, D.; Chen, X. Evaluation of the Agronomic Performance of Atrazine-Tolerant Transgenic japonica Rice Parental Lines for Utilization in Hybrid Seed Production. PLoS ONE 2014, 9, e108569. [Google Scholar] [CrossRef] [PubMed]
  2. Campanella, L.; Vitaliano, R. Atrazine toxicity reduction following H2O2/TiO2- photocatalyzed reaction and comparison with H2O2- photolytic reaction. Ann. Chim. 2007, 97, 123–134. [Google Scholar] [CrossRef]
  3. Héquet, V.; Le Cloirec, P.; Gonzalez, C.; Meunier, B. Photocatalytic degradation of atrazine by porphyrin and phthalocyanine complexes. Chemosphere 2000, 41, 379–386. [Google Scholar] [CrossRef] [PubMed]
  4. Ryberg, K.R.; Stone, W.W.; Baker, N.T. Causal factors for pesticide trends in streams of the United States: Atrazine and deethylatrazine. J. Environ. Qual. 2020, 49, 152–162. [Google Scholar] [CrossRef] [PubMed]
  5. Baldwin, A.K.; Corsi, S.R.; De Cicco, L.A.; Lenaker, P.L.; Lutz, M.A.; Sullivan, D.J.; Richards, K.D. Organic contaminants in Great Lakes tributaries: Prevalence and potential aquatic toxicity. Sci. Total Environ. 2016, 554–555, 42–52. [Google Scholar] [CrossRef] [PubMed]
  6. Murphy, M.; Hecker, M.; Coady, K.; Tompsett, A.; Jones, P.; Du Preez, L.; Everson, G.; Solomon, K.; Carr, J.; Smith, E.; et al. Atrazine concentrations, gonadal gross morphology and histology in ranid frogs collected in Michigan agricultural areas. Aquat. Toxicol. 2006, 76, 230–245. [Google Scholar] [CrossRef]
  7. Jiang, Z.; Li, J.; Jiang, D.; Gao, Y.; Chen, Y.; Wang, W.; Cao, B.; Tao, Y.; Wang, L.; Zhang, Y. Removal of atrazine by biochar-supported zero-valent iron catalyzed persulfate oxidation: Reactivity, radical production and transformation pathway. Environ. Res. 2020, 184, 109260. [Google Scholar] [CrossRef]
  8. Lang, D.H.; Rettie, A.E.; Böcker, R.H. Identification of enzymes involved in the metabolism of atrazine, terbuthylazine, ametryne, and terbutryne in human liver microsomes. Chem. Res. Toxicol. 1997, 10, 1037–1044. [Google Scholar] [CrossRef]
  9. Shamsedini, N.; Dehghani, M.; Nasseri, S.; Baghapour, M.A. Photocatalytic degradation of atrazine herbicide with Illuminated Fe+3-TiO2 Nanoparticles. J. Environ. Health Sci. Eng. 2017, 14, 15. [Google Scholar] [CrossRef]
  10. Supraja, P.; Singh, V.; Vanjari, S.R.K.; Govind Singh, S. Electrospun CNT embedded ZnO nanofiber-based biosensor for electrochemical detection of Atrazine: A step closure to single molecule detection. Microsyst. Nanoeng. 2020, 6, 3. [Google Scholar] [CrossRef]
  11. Wang, J.; Tafen, D.N.; Lewis, J.P.; Hong, Z.; Manivannan, A.; Zhi, M.; Li, M.; Wu, N. Origin of Photocatalytic Activity of Nitrogen-Doped TiO2 Nanobelts. J. Am. Chem. Soc. 2009, 131, 12290–12297. [Google Scholar] [CrossRef]
  12. Luo, Y.; Wang, J.; Yu, S.; Cao, Y.; Ma, K.; Pu, Y.; Zou, W.; Tang, C.; Gao, F.; Dong, L. Nonmetal element doped g-C3N4 with enhanced H2 evolution under visible light irradiation. J. Mater. Res. 2018, 33, 1268–1278. [Google Scholar] [CrossRef]
  13. He, H.; Huang, L.; Zhong, Z.; Tan, S. Constructing three-dimensional porous graphene-carbon quantum dots/g-C3N4 nanosheet aerogel metal-free photocatalyst with enhanced photocatalytic activity. Appl. Surf. Sci. 2018, 441, 285–294. [Google Scholar] [CrossRef]
  14. Li, P.; Liu, L.; An, W.; Wang, H.; Guo, H.; Liang, Y.; Cui, W. Ultrathin porous g-C3N4 nanosheets modified with AuCu alloy nanoparticles and C-C coupling photothermal catalytic reduction of CO2 to ethanol. Appl. Catal. B 2020, 266, 118618. [Google Scholar] [CrossRef]
  15. Deng, Y.; Tang, L.; Zeng, G.; Zhu, Z.; Yan, M.; Zhou, Y.; Wang, J.; Liu, Y.; Wang, J. Insight into highly efficient simultaneous photocatalytic removal of Cr(VI) and 2,4-diclorophenol under visible light irradiation by phosphorus doped porous ultrathin g-C3N4 nanosheets from aqueous media: Performance and reaction mechanism. Appl. Catal. B 2017, 203, 343–354. [Google Scholar] [CrossRef]
  16. Mohamed, M.A.; Zain, M.M.; Minggu, L.J.; Kassim, M.B.; Amin, N.A.S.; Salleh, W.W.; Salehmin, M.N.I.; Nasir, M.F.M.; Hir, Z.A.M. Constructing bio-templated 3D porous microtubular C-doped g-C3N4 with tunable band structure and enhanced charge carrier separation. Appl. Catal. B 2018, 236, 265–279. [Google Scholar] [CrossRef]
  17. Yang, Q.; Yang, C.C.; Lin, C.H.; Jiang, H.L. Metal-organic-framework-derived hollow N-doped porous carbon with ultrahigh concentrations of single Zn atoms for efficient carbon dioxide conversion. Angew. Chem. Int. Ed. 2019, 58, 3511–3515. [Google Scholar] [CrossRef]
  18. Tang, R.; Gong, D.; Deng, Y.; Xiong, S.; Deng, J.; Li, L.; Zhou, Z.; Zheng, J.; Su, L.; Yang, L. p-p stacked step-scheme PDI/g-C3N4/TiO2@Ti3C2 photocatalyst with enhanced visible photocatalytic degradation towards atrazine via peroxymonosulfate activation. Chem. Eng. J. 2022, 427, 131809. [Google Scholar] [CrossRef]
  19. Li, C.; Wu, H.; Zhu, D.; Zhou, T.; Yan, M.; Chen, G.; Sun, J.; Dai, G.; Ge, F.; Dong, H. High efficient charge separation driven directionally by pyridine rings grafted on carbon nitride edge for boosting photocatalytic hydrogen evolution. Appl. Catal. B 2021, 297, 120433. [Google Scholar] [CrossRef]
  20. Guo, F.; Chen, J.; Zhao, J.; Chen, Z.; Xia, D.; Zhan, Z.; Wang, Q. Z-scheme heterojunction g-C3N4@PDA/BiOBr with biomimetic polydopamine as electron transfer mediators for enhanced visible-light driven degradation of sulfamethoxazole. Chem. Eng. J. 2020, 386, 124014. [Google Scholar] [CrossRef]
  21. Tang, R.; Gong, D.; Deng, Y.; Xiong, S.; Zheng, J.; Li, L.; Zhou, Z.; Su, L.; Zhao, J. p-p stacking derived from graphene-like biochar/g-C3N4 with tunable band structure for photocatalytic antibiotics degradation via peroxymonosulfate activation. J. Hazard. Mater. 2021, 423, 126944. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, G.; Huang, R.; Zhang, J.; Mao, J.; Wang, D.; Li, Y. Synergistic modulation of the separation of photo-generated carries via engineering of dual atomic sites for promoting photocatalytic performance. Adv. Mater. 2021, 33, e2105904. [Google Scholar] [CrossRef] [PubMed]
  23. Ma, P.; Zhang, X.; Wang, C.; Wang, Z.; Wang, K.; Feng, Y.; Wang, J.; Zhai, Y.; Deng, J.; Wang, L.; et al. Band alignment of homojunction by anchoring CN quantum dots on g-C3N4 (0D/2D) enhance photocatalytic hydrogen peroxide evolution. Appl. Catal. B 2021, 298, 120736. [Google Scholar] [CrossRef]
  24. Zhen, W.; Sun, J.; Ning, X.; Shi, X.; Xue, C. Polymeric carbon nitride with internal n-p homojunctions for efficient photocatalytic CO2 reduction coupled with cyclohexene oxidation. Appl. Catal. B 2021, 298, 120568. [Google Scholar] [CrossRef]
  25. Li, Y.; Fang, Y.; Cao, Z.; Li, N.; Chen, D.; Xu, Q.; Lu, J. Construction of g-C3N4/ PDI@MOF heterojunctions for the highly efficient visible light-driven degradation of pharmaceutical and phenolic micropollutants. Appl. Catal. B 2019, 250, 150–162. [Google Scholar] [CrossRef]
  26. Zhu, X.; Yang, J.; Zhu, X.; Yuan, J.; Zhou, M.; She, X.; Yu, Q.; Song, Y.; She, Y.; Hua, Y.; et al. Exploring deep effects of atomic vacancies on activating CO2 photoreduction via rationally designing indium oxide photocatalysts. Chem. Eng. J. 2021, 422, 129888. [Google Scholar] [CrossRef]
  27. Zhu, X.; Zhou, G.; Yi, J.; Ding, P.; Yang, J.; Zhong, K.; Song, Y.; Hua, Y.; Zhu, X.; Yuan, J.; et al. Accelerated Photoreduction of CO2 to CO over a Stable Heterostructure with a Seamless Interface. ACS Appl. Mater. Interfaces 2021, 13, 39523–39532. [Google Scholar] [CrossRef]
  28. Castro Neto, A.H.; Guinea, F.; Peres, N.M.R.; Novoselov, K.S.; Geim, A.K. The electronic properties of graphene. Rev. Mod. Phys. 2009, 81, 109–162. [Google Scholar] [CrossRef]
  29. Du, X.; Skachko, I.; Barker, A.; Andrei, E.Y. Approaching ballistic transport in suspended graphene. Nat. Nanotechnol. 2008, 3, 491–495. [Google Scholar] [CrossRef]
  30. Huang, Y.; Liang, J.; Chen, Y. An overview of the applications of graphene-based materials in supercapacitors. Small 2012, 8, 1805–1834. [Google Scholar] [CrossRef]
  31. Lee, J.S.; You, K.H.; Park, C.B. Highly Photoactive, Low Bandgap TiO2 Nanoparticles Wrapped by Graphene. Adv. Mater. 2012, 24, 1084–1088. [Google Scholar] [CrossRef] [PubMed]
  32. Stoller, M.D.; Park, S.; Zhu, Y.; An, J.; Ruoff, R.S. Graphene-Based Ultracapacitors. Nano Lett. 2008, 8, 3498–3502. [Google Scholar] [CrossRef] [PubMed]
  33. Burghard, M.; Klauk, H.; Kern, K. Carbon-Based Field-Effect Transistors for Nanoelectronics. Adv. Mater. 2009, 21, 2586–2600. [Google Scholar] [CrossRef]
  34. Fujita, M.; Wakabayashi, K.; Nakada, K.; Kusakabe, K. Peculiar Localized State at Zigzag Graphite Edge. J. Phys. Soc. Jpn. 1996, 65, 1920–1923. [Google Scholar] [CrossRef]
  35. Kuilla, T.; Bhadra, S.; Yao, D.; Kim, N.H.; Bose, S.; Lee, J.H. Recent advances in graphene based polymer composites. Prog. Polym. Sci. 2010, 35, 1350–1375. [Google Scholar] [CrossRef]
  36. Niyogi, S.; Bekyarova, E.; Itkis, M.E.; McWilliams, J.L.; Hamon, M.A.; Haddon, R.C. Solution Properties of Graphite and Graphene. J. Am. Chem. Soc. 2006, 128, 7720–7721. [Google Scholar] [CrossRef]
  37. Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J.W.; Potts, J.R.; Ruoff, R.S. Graphene and Graphene Oxide: Synthesis, Properties, and Applications. Adv. Mater. 2010, 22, 3906–3924. [Google Scholar] [CrossRef]
  38. Hemidouche, S.; Boudriche, L.; Boudjemaa, A.; Hamoudi, S. Removal of lead (II) and cadmium (II) cations from water using surface-modified graphene. Can. J. Chem. Eng. 2017, 95, 508–515. [Google Scholar] [CrossRef]
  39. Pei, S.; Cheng, H.-M. The reduction of graphene oxide. Carbon 2012, 50, 3210–3228. [Google Scholar] [CrossRef]
  40. Renteria, J.D.; Ramirez, S.; Malekpour, H.; Alonso, B.; Centeno, A.; Zurutuza, A.; Cocemasov, A.I.; Nika, D.L.; Balandin, A.A. Strongly Anisotropic Thermal Conductivity of Free-Standing Reduced Graphene Oxide Films Annealed at High Temperature. Adv. Funct. Mater. 2015, 25, 4664–4672. [Google Scholar] [CrossRef]
  41. Novoselov, K.S.; Fal′ko, V.I.; Colombo, L.; Gellert, P.R.; Schwab, M.G.; Kim, K. A roadmap for graphene. Nature 2012, 490, 192–200. [Google Scholar] [CrossRef] [PubMed]
  42. Park, S.; Ruoff, R.S. Chemical methods for the production of graphenes. Nat. Nanotechnol. 2009, 4, 217–224. [Google Scholar] [CrossRef] [PubMed]
  43. Xu, B.; Ahmed, M.B.; Zhou, J.L.; Altaee, A.; Xu, G.; Wu, M. Graphitic carbon nitride based nanocomposites for the photocatalysis of organic contaminants under visible irradiation: Progress, limitations and future directions. Sci. Total Environ. 2018, 633, 546–559. [Google Scholar] [CrossRef]
  44. Luo, Y.; Zhu, Y.; Han, Y.; Ye, H.; Liu, R.; Lan, Y.; Xue, M.; Xie, X.; Yu, S.; Zhang, L.; et al. g-C3N4-based photocatalysts for organic pollutant removal: A critical review. Carbon Res. 2023, 2, 14. [Google Scholar] [CrossRef]
  45. Mohamed, R.M. Synthesis and characterization of AgCl@graphitic carbon nitride hybrid materials for the photocatalytic degradation of atrazine. Ceram. Int. 2015, 41, 1197–1204. [Google Scholar] [CrossRef]
  46. Jo, W.K.; Lee, J.Y.; Selvam, N.C.S. Synthesis of MoS2 nanosheets loaded ZnO-g-C3N4 nanocomposites for enhanced photocatalytic applications. Chem. Eng. J. 2016, 289, 306–318. [Google Scholar] [CrossRef]
  47. Jo, W.K.; Selvam, N.C.S. Z-scheme CdS/g-C3N4 composites with RGO as an electron mediator for efficient photocatalytic H2 production and pollutant degradation. Chem. Eng. J. 2017, 317, 913–924. [Google Scholar] [CrossRef]
  48. Dikdim, J.M.D.; Gong, Y.; Noumi, G.B.; Sieliechi, J.M.; Zhao, X.; Ma, N.; Yang, M.; Tchatchueng, J.B. Peroxymonosulfate improved photocatalytic degradation of atrazine by activated carbon/graphitic carbon nitride composite under visible light irradiation. Chemosphere 2019, 217, 833–842. [Google Scholar] [CrossRef]
  49. Wang, X.; Meng, J.; Zhang, X.; Liu, Y.; Ren, M.; Yang, Y.; Guo, Y. Controllable Approach to Carbon-Deficient and Oxygen-Doped Graphitic Carbon Nitride: Robust Photocatalyst Against Recalcitrant Organic Pollutants and the Mechanism Insight. Adv. Funct. Mater. 2021, 31, 2010763. [Google Scholar] [CrossRef]
  50. Fina, F.; Callear, S.K.; Carins, G.M.; Irvine, J.T.S. Structural Investigation of Graphitic Carbon Nitride via XRD and Neutron Diffraction. Chem. Mater. 2015, 27, 2612–2618. [Google Scholar] [CrossRef]
  51. Wang, Z.; Huo, Y.; Fan, Y.; Wu, R.; Wu, H.; Wang, F.; Xu, X. Facile synthesis of carbon-rich g-C3N4 by copolymerization of urea and tetracyanoethylene for photocatalytic degradation of Orange II. J. Photochem. Photobiol. A Chem. 2018, 358, 61–69. [Google Scholar] [CrossRef]
  52. Xiang, Q.; Yu, J.; Jaroniec, M. Preparation and Enhanced Visible-Light Photocatalytic H2-Production Activity of Graphene/C3N4 Composites. J. Phys. Chem. C 2011, 115, 7355–7363. [Google Scholar] [CrossRef]
  53. Zhang, R.; Zhang, X.; Liu, S.; Tong, J.; Kong, F.; Sun, N.; Han, X.; Zhang, Y. Enhanced photocatalytic activity and optical response mechanism of porous graphitic carbon nitride (g-C3N4) nanosheets. Mater. Res. Bull. 2021, 140, 111263. [Google Scholar] [CrossRef]
  54. Ma, T.; Dai, S.; Jaroniec, M.; Qiao, S.Z. Graphitic Carbon Nitride Nanosheet–Carbon Nanotube Three-Dimensional Porous Composites as High-Performance Oxygen Evolution Electrocatalysts. Angew. Chem. Int. Ed. 2014, 53, 7281–7285. [Google Scholar] [CrossRef]
  55. Miao, X.; Ji, Z.; Wu, J.; Shen, X.; Wang, J.; Kong, L.; Liu, M.; Song, C. g-C3N4/AgBr nanocomposite decorated with carbon dots as a highly efficient visible-light-driven photocatalyst. J. Colloid Interface Sci. 2017, 502, 24–32. [Google Scholar] [CrossRef]
  56. Yuan, B.; Wei, J.; Hu, T.; Yao, H.; Jiang, Z.; Fang, Z.; Chu, Z. Simple synthesis of g-C3N4/rGO hybrid catalyst for the photocatalytic degradation of rhodamine B. Chin. J. Catal. 2015, 36, 1009–1016. [Google Scholar] [CrossRef]
  57. Sengupta, I.; Chakraborty, S.; Talukdar, M.; Pal, S.K.; Chakraborty, S. Thermal reduction of graphene oxide: How temperature influences purity. J. Mater. Res. 2018, 33, 4113–4122. [Google Scholar] [CrossRef]
  58. Yang, X.; Tian, Z.; Chen, Y.; Huang, H.; Hu, J. One-pot calcination preparation of graphene/g–C3N4–Co photocatalysts with enhanced visible light photocatalytic activity. Int. J. Hydrog. Energy 2020, 45, 12889–12902. [Google Scholar] [CrossRef]
  59. Lin, Z.; Wang, X. Nanostructure Engineering and Doping of Conjugated Carbon Nitride Semiconductors for Hydrogen Photosynthesis. Angew. Chem. Int. Ed. 2013, 52, 1735–1738. [Google Scholar] [CrossRef]
  60. Xu, T.; Zhang, L.; Cheng, H.; Zhu, Y. Significantly enhanced photocatalytic performance of ZnO via graphene hybridization and the mechanism study. Appl. Catal. B Environ. 2011, 101, 382–387. [Google Scholar] [CrossRef]
  61. Zhang, Y.; Mori, T.; Niu, L.; Ye, J. Non-covalent doping of graphitic carbon nitride polymer with graphene: Controlled electronic structure and enhanced optoelectronic conversion. Energy Environ. Sci. 2011, 4, 4517–4521. [Google Scholar] [CrossRef]
  62. Negro, P.; Cesano, F.; Casassa, S.; Scarano, D. Combined DFT-D3 Computational and Experimental Studies on g-C3N4: New Insight into Structure, Optical, and Vibrational Properties. Mater. Internet 2023, 16, 3644. [Google Scholar] [CrossRef]
  63. Xu, C.; Liu, X.; Li, D.; Chen, Z.; Yang, J.; Huang, J.; Pan, H. Coordination of π-Delocalization in g-C3N4 for Efficient Photocatalytic Hydrogen Evolution under Visible Light. ACS Appl. Mater. Interfaces 2021, 13, 20114–20124. [Google Scholar] [CrossRef]
  64. Li, Y.; Jiang, Y.; Ruan, Z.; Lin, K.; Yu, Z.; Zheng, Z.; Xua, X.; Yuan, Y. Simulation-guided synthesis of graphitic carbon nitride beads with 3D interconnected and continuous meso/macropore channels for enhanced light absorption and photocatalytic performance. J. Mater. Chem. A 2017, 5, 21300–21312. [Google Scholar] [CrossRef]
  65. Dong, G.; Zhang, Y.; Pan, Q.; Qiu, J. A fantastic graphitic carbon nitride (g-C3N4) material: Electronic structure, photocatalytic and photoelectronic properties. J. Photochem. Photobiol. C Photochem. Rev. 2014, 20, 33–50. [Google Scholar] [CrossRef]
  66. Ma, J.; Tan, X.; Jiang, F.; Yu, T. Graphitic C3N4 nanosheet-sensitized brookite TiO2 to achieve photocatalytic hydrogen evolution under visible light. Catal. Sci. Technol. 2017, 7, 3275–3282. [Google Scholar] [CrossRef]
  67. Du, A.; Sanvito, S.; Li, Z.; Wang, D.; Jiao, Y.; Liao, T.; Smith, S.C. Hybrid Graphene and Graphitic Carbon Nitride Nanocomposite: Gap Opening, Electron–Hole Puddle, Interfacial Charge Transfer, and Enhanced Visible Light Response. J. Am. Chem. Soc. 2012, 134, 4393–4397. [Google Scholar] [CrossRef]
  68. Chakrabarti, S.; Dutta, B.K. Photocatalytic degradation of model textile dyes in wastewater using ZnO as semiconductor catalyst. J. Hazard. Mater. 2004, 112, 269–278. [Google Scholar] [CrossRef]
  69. Huang, M.; Xu, C.; Wu, Z.; Huang, Y.; Lin, J.; Wu, J. Photocatalytic discolorization of methyl orange solution by Pt modified TiO2 loaded on natural zeolite. Dye. Pigment. 2008, 77, 327–334. [Google Scholar] [CrossRef]
  70. Sharma, B.; Boruah, P.K.; Yadav, A.; Das, M.R. TiO2–Fe2O3 nanocomposite heterojunction for superior charge separation and the photocatalytic inactivation of pathogenic bacteria in water under direct sunlight irradiation. J. Environ. Chem. Eng. 2018, 6, 134–145. [Google Scholar] [CrossRef]
  71. Dehghani, M.; Haidari, E.; Shahsavani, S.; Shamsedini, N. Removal of Nitrate in the Aqueous Phase Using Granular Ferric Hydroxide. Jundishapur J. Health Sci. 2015, 7, e26419. [Google Scholar] [CrossRef]
  72. Xu, L.; Zang, H.; Zhang, Q.; Chen, Y.; Wei, Y.; Yan, J.; Zhao, Y. Photocatalytic degradation of atrazine by H3PW12O40/Ag–TiO2: Kinetics, mechanism and degradation pathways. Chem. Eng. J. 2013, 232, 174–182. [Google Scholar] [CrossRef]
  73. Phonsy, P.D.; Anju, S.G.; Jyothi, K.P.; Yesodharan, S.; Yesodharan, E.P. Semiconductor Mediated Photocatalytic Degradation of Plastics and Recalcitrant Organic Pollutants in Water: Effect of Additives and Fate of Insitu Formed H2O2. J. Adv. Oxid. Technol. 2015, 18, 85–97. [Google Scholar] [CrossRef]
  74. Xue, Y.; Wang, P.; Wang, C.; Ao, Y. Efficient degradation of atrazine by BiOBr/UiO-66 composite photocatalyst under visible light irradiation: Environmental factors, mechanisms and degradation pathways. Chemosphere 2018, 203, 497–505. [Google Scholar] [CrossRef]
  75. Tan, L.-L.; Chai, S.-P.; Mohamed, A.R. Synthesis and Applications of Graphene-Based TiO2 Photocatalysts. ChemSusChem 2012, 5, 1868–1882. [Google Scholar] [CrossRef]
  76. Hartley, D.; Kidd, H. The Agrochemicals Handbook, 3rd ed.; Royal Society of Chemistry: Cambridge, UK, 1991. [Google Scholar]
  77. Liu, X.; Zong, H.; Tan, X.; Wang, X.; Qiu, J.; Kong, F.; Zhang, J.; Fang, S. Facile synthesis of modified carbon nitride with enhanced activity for photocatalytic degradation of atrazine. J. Environ. Chem. Eng. 2021, 9, 105807. [Google Scholar] [CrossRef]
  78. Yamaguchi, N.U.; Bergamasco, R.; Hamoudi, S. Magnetic MnFe2O4–graphene hybrid composite for efficient removal of glyphosate from water. Chem. Eng. J. 2016, 295, 391–402. [Google Scholar] [CrossRef]
  79. Zhu, K.; Du, Y.; Liu, J.; Fan, X.; Duan, Z.; Song, G.; Meng, A.; Li, Z.; Li, Q. Graphitic Carbon Nitride (g-C3N4) Nanosheets/Graphene Composites: In Situ Synthesis and Enhanced Photocatalytic Performance. J. Nanosci. Nanotechnol. 2017, 17, 2515–2519. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of pure g-C3N4 and graphene/g-C3N4 nanocomposites; a.u. stands for arbitrary units.
Figure 1. XRD patterns of pure g-C3N4 and graphene/g-C3N4 nanocomposites; a.u. stands for arbitrary units.
Catalysts 13 01265 g001
Figure 2. SEM images of g-C3N4 (a), Graphene/g-C3N4 (b) and TEM images of g-C3N4 (c), Graphene/g-C3N4 (d).
Figure 2. SEM images of g-C3N4 (a), Graphene/g-C3N4 (b) and TEM images of g-C3N4 (c), Graphene/g-C3N4 (d).
Catalysts 13 01265 g002
Figure 3. FTIR spectra of g-C3N4 and graphene/g-C3N4 materials; a.u. stands for arbitrary units.
Figure 3. FTIR spectra of g-C3N4 and graphene/g-C3N4 materials; a.u. stands for arbitrary units.
Catalysts 13 01265 g003
Figure 4. Nitrogen physisorption isotherms of photocatalysts.
Figure 4. Nitrogen physisorption isotherms of photocatalysts.
Catalysts 13 01265 g004
Figure 5. (a) Photoluminescence spectra of g-C3N4 and graphene/g-C3N4 under light (excitation, 330 nm), (b) photoluminescence intensity spectra of g-C3N4 and graphene/g-C3N4 versus time; a.u. stands for arbitrary units.
Figure 5. (a) Photoluminescence spectra of g-C3N4 and graphene/g-C3N4 under light (excitation, 330 nm), (b) photoluminescence intensity spectra of g-C3N4 and graphene/g-C3N4 versus time; a.u. stands for arbitrary units.
Catalysts 13 01265 g005
Figure 6. UV-vis absorption spectra (a) and Tauc plots (b) for g-C3N4 and Graphene/g-C3N4; a.u. stands for arbitrary units.
Figure 6. UV-vis absorption spectra (a) and Tauc plots (b) for g-C3N4 and Graphene/g-C3N4; a.u. stands for arbitrary units.
Catalysts 13 01265 g006
Figure 7. Atrazine adsorption in the dark and the time profile of atrazine photodegradation under visible light irradiation in the presence of g-C3N4 and graphene/g-C3N4. Photocatalyst loading: 0.4 g/L; atrazine initial concentration: 10 mg/L; temperature: 25 °C. Error bars refer to standard deviation.
Figure 7. Atrazine adsorption in the dark and the time profile of atrazine photodegradation under visible light irradiation in the presence of g-C3N4 and graphene/g-C3N4. Photocatalyst loading: 0.4 g/L; atrazine initial concentration: 10 mg/L; temperature: 25 °C. Error bars refer to standard deviation.
Catalysts 13 01265 g007
Figure 8. (a) Effect of graphene/g-C3N4 photocatalyst loading on atrazine adsorption in the dark and the time profile of atrazine photodegradation under visible light irradiation, (b) Atrazine dimensionless residual concentration at t = 300 min as a function of photocatalyst loading. Atrazine initial concentration: 10 mg/L; temperature: 25 °C. Error bars refer to standard deviation.
Figure 8. (a) Effect of graphene/g-C3N4 photocatalyst loading on atrazine adsorption in the dark and the time profile of atrazine photodegradation under visible light irradiation, (b) Atrazine dimensionless residual concentration at t = 300 min as a function of photocatalyst loading. Atrazine initial concentration: 10 mg/L; temperature: 25 °C. Error bars refer to standard deviation.
Catalysts 13 01265 g008
Figure 9. Effect of suspension pH on atrazine adsorption in the dark and the time profile of atrazine photodegradation under visible light irradiation over graphene/g-C3N4. Photocatalyst loading: 0.4 g/L; atrazine initial concentration: 10 mg/L; temperature: 25 °C. Error bars refer to standard deviation.
Figure 9. Effect of suspension pH on atrazine adsorption in the dark and the time profile of atrazine photodegradation under visible light irradiation over graphene/g-C3N4. Photocatalyst loading: 0.4 g/L; atrazine initial concentration: 10 mg/L; temperature: 25 °C. Error bars refer to standard deviation.
Catalysts 13 01265 g009
Figure 10. A schematic illustration of charge transfer and separation in graphene/g-C3N4 composite for photocatalytic degradation of atrazine under visible light irradiation.
Figure 10. A schematic illustration of charge transfer and separation in graphene/g-C3N4 composite for photocatalytic degradation of atrazine under visible light irradiation.
Catalysts 13 01265 g010
Figure 11. Stability and reusability of graphene/g-C3N4. Photocatalyst loading: 0.4 g/L; atrazine initial concentration: 10 mg/L; temperature: 25 °C.
Figure 11. Stability and reusability of graphene/g-C3N4. Photocatalyst loading: 0.4 g/L; atrazine initial concentration: 10 mg/L; temperature: 25 °C.
Catalysts 13 01265 g011
Table 1. Different photocatalysts performances in the degradation of aqueous atrazine under visible light at ambient temperature.
Table 1. Different photocatalysts performances in the degradation of aqueous atrazine under visible light at ambient temperature.
Initial Concentration
(mg/L)
Photocatalyst
(Loading g/L)
Time
(min)
Degradation
(%)
Ref.
100g-C3N4 (0.4)6010[45]
100AgCl/g-C3N4 (0.4)60100[45]
10ZnO–g-C3N4/MoS2 (0.3)30080[46]
10CdS/RGO/g-C3N4 a (0.2)30090[47]
10AC/g-C3N4 b (1)12090[48]
10Vc OCN c (1)120100[49]
1TEA-CN d (0.3)6090[77]
10PDI-g-C3N4/g-C3N4 e (0.8)6075[18]
10Graphene/g-C3N4 (0.3)300100This work
a RGO: reduced graphene oxide. b AC: Activated carbon. c Vc-OCN: Carbon vacancy-oxygen doped graphitic carbon nitride. d TEA-CN: Triethanolamine modified carbon nitride. e PDI-g-C3N4/g-C3N4: pyromellitic diimide modified g-C3N4/g-C3N4 homojunction.
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

Altendji, K.; Hamoudi, S. Efficient Photocatalytic Degradation of Aqueous Atrazine over Graphene-Promoted g-C3N4 Nanosheets. Catalysts 2023, 13, 1265. https://doi.org/10.3390/catal13091265

AMA Style

Altendji K, Hamoudi S. Efficient Photocatalytic Degradation of Aqueous Atrazine over Graphene-Promoted g-C3N4 Nanosheets. Catalysts. 2023; 13(9):1265. https://doi.org/10.3390/catal13091265

Chicago/Turabian Style

Altendji, Khaoula, and Safia Hamoudi. 2023. "Efficient Photocatalytic Degradation of Aqueous Atrazine over Graphene-Promoted g-C3N4 Nanosheets" Catalysts 13, no. 9: 1265. https://doi.org/10.3390/catal13091265

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