Next Article in Journal
Constructing Active Sites on Self-Supporting Ti3C2Tx (T = OH) Nanosheets for Enhanced Photocatalytic CO2 Reduction into Alcohols
Next Article in Special Issue
Synergistic Effect of Amorphous Ti(IV)-Hole and Ni(II)-Electron Cocatalysts for Enhanced Photocatalytic Performance of Bi2WO6
Previous Article in Journal
Selenium Oxoanions Removal from Wastewater by MoS42− Intercalated FeMgAl LDH: Catalytic Roles of Fe and Mechanism Insights
Previous Article in Special Issue
Efficient Visible-Light Driven Photocatalytic Hydrogen Production by Z-Scheme ZnWO4/Mn0.5Cd0.5S Nanocomposite without Precious Metal Cocatalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Z-Scheme g-C3N5/Cu3TiO4 Heterojunction Enhanced Photocatalytic Performance of Chromene-3-Carbonitriles Synthesis under Visible Light Irradiation

by
Murugan Arunachalapandi
,
Thangapandi Chellapandi
,
Gunabalan Madhumitha
,
Ravichandran Manjupriya
,
Kumar Aravindraj
and
Selvaraj Mohana Roopan
*
Chemistry of Heterocycles & Natural Product Research Laboratory, Department of Chemistry, School of Advanced Sciences, Vellore Institute of Technology, Vellore 632 014, Tamil Nadu, India
*
Author to whom correspondence should be addressed.
Submission received: 1 November 2022 / Revised: 28 November 2022 / Accepted: 2 December 2022 / Published: 6 December 2022
(This article belongs to the Special Issue Advances in Heterojunction Photocatalysts)

Abstract

:
In order to make the synthesis of pharmaceutically active carbonitriles efficient, environmentally friendly, and sustainable, the method is regularly examined. Here, we introduce a brand-new, very effective Cu3TiO4/g-C3N5 photocatalyst for the production of compounds containing chromene-3-carbonitriles. The direct Z-Scheme photo-generated charge transfer mechanism used by the Cu3TiO4/g-C3N5 photocatalyst results in a suppressed rate of electron-hole pair recombination and an increase in photocatalytic activity. Experiments showed that the current method has some advantages, such as using an environmentally friendly and sustainable photocatalyst, having a simple procedure, quick reaction times, a good product yield (82–94%), and being able to reuse the photocatalyst multiple times in a row without noticeably decreasing its photocatalytic performance.

Graphical Abstract

1. Introduction

Multi-component reactions (MCRs) have garnered a lot of interest recently as a method for creating a variety of heterocyclic compounds with applicability in medicine, pharmaceuticals, and agrochemistry [1,2]. One-pot MCRs, which induce organic reactants to react together in a single step, have emerged as an alternative platform for organic chemists due to their simple operation, effective purification techniques, side products, and quick turnaround times [3,4]. In simple terms, a multi-component reaction is one in which three or more reactants mix in a single vessel to create a single output that effectively contains every atom of the starting elements (with the exception of condensation products, such as water molecules, hydrochloric acid, or methanol) [5,6,7].
A number of naturally occurring O-bearing heterocyclic organic moieties, which are widely present in edible fruits and vegetables, contain the chromene and related structural skeletons as their major structural motifs [8,9]. Numerous industries, like agrochemistry, medicine, as well as material sciences, use carbonitrile-based heterocycles and their derivatives as ligands and catalysts [10]. Carbonitrile compounds have a high level of therapeutic activity and can be employed as antibacterial [11,12], anti-tumor [13,14], anti-inflammatory, anti-cancer [15,16,17], and anti-convulsant drugs [18]. Compounds with a carbonitrile base are also employed to make electrical and optical materials as well as anion sensors [19,20]. Given the many different uses for chromenes, synthetic chemists have been exploring new and effective catalysts. In the traditional way of making carbonitrile-based chemicals, the product needed to be made at a high temperature and through a long process called “synthesis”.
As a result of growing environmental worry, a lot of work has been put into creating innovative procedures that reduce emissions in chemical synthesis. From the perspective of green and sustainable chemistry [21,22,23], designing effective and affordable chemical processes utilising heterogeneous catalysts to create fine chemicals and pharmaceutical goods through MCRs has recently attracted a lot of attention from academics and businesses. The fundamental drawback of heterogeneous catalysts is, however, their decreased catalytic activity. The best way to deal with this issue is to minimize the size of the particles in the heterogeneous catalyst [24,25]. Metal oxide nanocatalysts combine the benefits of photo and heterogeneous catalysts. Although this type of material has many valuable advantages, including a high surface area, efficient particle distribution, a straightforward recovery procedure, effective light absorption property, good stability, and non-toxic behaviour, their propensity to agglomerate and oxidise is a drawback [26,27,28]. Immobilizing organic molecules on metal oxide nanoparticles is thought to be a workable way to address this drawback.
Given the significance of metal oxide, nanoparticles (metal oxide NPs) have received a lot of attention and are widely employed in water splitting, catalysis, photocatalysis, energy, and environmental remediation [29,30,31,32]. Titanium dioxide (TiO2) and copper oxide (CuO) nanoparticles have attracted a great deal of interest. It was discovered that CuO and TiO2 NPs worked effectively as catalysts for the oxidation of organic molecules as well as CO and NO [33,34]. Additionally, in last decade, a lot of attention has been focused on the use of CuO and TiO2 NPs for chemical reactions [35]. CuO NPs’ high levels of electron-hole pair recombination and TiO2’s less obvious active nature are some of its disadvantages. We created the Cu3TiO4 NPs for the organic reaction based on the aforementioned investigative information. In the context of heterogeneous nanocatalysts, this class of catalysts seems to be one of the most promising approaches toward effective reactions under mild and ecologically nonthreatening circumstances.
Organic synthesis research in the area of photocatalysis, notably with regard to visible light, is appealing [36]. Heterogeneous photocatalysis has been given a significant role by carbon-based materials. Due to a variety of characteristics, including stability, narrow bandgap, 2D nanomaterial, excellent visible-light active absorption, low density, and low toxicity, graphitic-carbon nitride (g-C3N4) is an effective catalyst under visible light [37,38]. Additionally, its layered structure makes the anchor metal an appropriate substrate for its direct use as a heterogeneous photocatalyst. A variety of modified carbon nitrides, including nanosheets, nanorods, metal-doped and non-metal modified carbon nitrides are effective nanomaterials [39,40,41]. An intriguing method has been recently created to modify the polymeric structure of C3N4 in order to create novel carbon nitrides with a wider visible-light sensitive range. For instance, integrating an N-rich 3-amino-1,2,4-triazole unit into the s-heptazine motif led to the highly ordered C3N5, which can increase the C:N atom ratio from 3:4–3:5. The lower bandgap energy is a result of the extra N-atoms connected to the terminal NH2 groups in the form of 1,2,4-triazole, which are different from the usual C3N4 in that they are bridged by a N-atom through sp2 hybridization [42,43,44]. The construction of heterojunction photocatalysts based on C3N4 has received extensive research, due to its viability and efficiency for the spatial separation of photogenerated electron-hole pairs. Appropriate catalyst materials can generate heterojunctions, which can lower the activation energy needed for the reduction reaction. The secondary target is to encourage the division of photogenerated charge carriers while simultaneously providing more active locations for the photocatalytic activity. Meticulous design of a heterojunction is a promising method to increase photocatalytic activity, because of its rapid charge carrier separation caused by the integrated electric field [45,46,47,48]. A switch from photocatalysts based on C3N5 to those based on the organic transformation reaction is driving recent advancements for cleaner, more sustainable chemistry. The heterojunction is followed by carbon nitride-based catalysts like Type-II and Z-Scheme. A Z-scheme system has also been built with a pertinent shuttle mediator to obtain better light-harvesting potential with high charge separation performance and sturdy redox ability, furthermore to photocatalysts with conventional heterojunctions that can successfully separate photogenerated electron-hole pairs via band alignment [49,50]. Numerous studies on g-C3N4 linked to the Z-Scheme electron transfer mechanism have been reported, including those on TiO2/g-C3N4 [51], CuO/g-C3N4 [52], ZnO/g-C3N4 [53], WO3/g-C3N4 [54], V2O5/C3N4 [55], MnO2/C3N4, [56] and Bi2O3/g-C3N4 [57], among others. In recent years, g-C3N5-based heterojunction production has become increasingly significant in photocatalysis and other applications. There are currently very few studies on Z-Scheme mechanisms for g-C3N5-based heterojunctions, including Ag3PO4/C3N5 [58], FeOCl/g-C3N5 [59], Bi4O5I2/g-C3N5 [60], LaCoO3/g-C3N5 [61] and Bi2WO6/g-C3N5 [62,63]. Still, there is no report carried out on Cu3TiO4/g-C3N5 heterojunction fabrication and photocatalytic application.
Light-emitting diodes (LEDs) are the best type of light source for photocatalytic applications. BLUE LEDs have benefits over other light sources in terms of efficiency, the qualities of power, compatibility, longevity, and environmental friendliness. Due to their unique characteristics, BLUE LEDs offer an improvement over traditional lighting sources and allow for more innovative scope when designing different photochemical reactions [64].
In this research, we have developed a novel and very effective Cu3TiO4/g-C3N5 photocatalyst for the construction of chromene-3-carbonitrile molecules under BLUE LED light. The direct Z-scheme charge separation mechanism employed by the Cu3TiO4/g-C3N5 hetero-junction photocatalyst reduces the rate of electron-hole (e/h+) pair recombination of Cu3TiO4 and increases its photocatalytic activity. The as-prepared g-C3N5/Cu3TiO4 nanocomposites showed better photocatalytic activity and stability when compared to pure Cu3TiO4 and pure C3N5. The spatial separation of photoinduced charge carriers could be attributed to the improved photocatalytic performance. The characteristics of the obtained catalysts have been studied with the help of XRD, UV-Vis DRS, FT-IR, FE-SEM, TEM, Zetapotential, and XPS. The applications for the catalyst have also been demonstrated and characterized. This work will be opening up a new path in the field of photocatalysts and MCRs and can further be developed for a wide range of applications other than the proposed one.

2. Results and Discussion

2.1. Characterization of Prepared Nanomaterials

The synthesized nanocomposite nanomaterial crystalline phase structure was investigated using the X-ray diffraction technique. In Figure 1, the pristine g-C3N5 sheets peak at 13.1° was related to (100), and another strong peak at 27.2° was related to (002) for graphitic g-C3N5 sheets (JCPDS card-87-1526) [65]. Peaks for Cu3TiO4 nanoparticles were found at 2theta = 26.4, 34.2, 38.6, 51.5, 59.1, 63.2, and 70.3° in relation to planes (004) (101) (102) (104) (006) (110) (112) (107), which corresponds well to JCPDS card no. 83-1285 [66]. Both are well-reflected in all CN5CT nanocomposites. Figure 1 shows the p-XRD patterns obtained for g-C3N5/Cu3TiO4 with varying wt. % Cu3TiO4. The XRD patterns of the 5% CN5CT composite showed no discernible change, while less Cu3TiO4 has been deducted because of the low loading percentage. The intensity of Cu3TiO4 nanoparticles peaks increased as the Cu3TiO4 nanoparticles ratio was increased to 10%, 20%, and 30%. Finally, Cu3TiO4 nanoparticles were supported on g-C3N5 nanosheets. As confirmed, they have shown good crystalline nature.
In BET analysis the surface area of the designed nanocomposites was examined, and depicted in Figure 2a. The surface area of Cu3TiO4 nanoparticles, g-C3N5 and g-C3N4 values are 10, 15, and 14 m2/g respectively. Then 10% CN5CT nanocomposite has 19 m2/g after supporting Cu3TiO4-nanoparticles, which thereby increased the surface area. The Cu3TiO4/g-C3N4 has less surface area compared to 10% CN5CT, as shown in Figure 2a. From BET analysis, 10%CN5CT exhibited more surface and active sites compared to others. In FT-IR spectroscopy (Figure 2b), g-C3N5 sheets showed peaks at 808, 1234, 1405, 1575, and 2160 cm−1, respectively. The signal at 808 cm−1 corresponds to the existence of triazine units in g-C3N5 nanosheets. NH2 and NH peaks were obtained at 3500 cm−1 and triazine unit stretching bands were found at 1200 cm−1. The g-C3N5 nanosheet FT-IR report is consistent with previous literature. [67,68]. The FT-IR spectrum of Cu3TiO4 exhibited a range between 400–700 cm−1 bands, which is owing to Ti-O-Ti and Cu-O stretching vibration [69]. After that, Cu3TiO4 nanoparticles supported the slight shift that happened in 5%, 10%, 20%, and 30% CN5CT nanocomposite.
In XPS, the chemical state of the prepared 10% CN5CT was exposed in Figure S1. Figure S1a shows the survey spectrum of the synthesized 10% CN5CT nanocomposite and all the signals confirmed the presence of C, N, Cu, Ti, and O. For C 1s, in Figure S1b, the three strong peaks at ~287, 285.8 and 283 eV show the presence of C 1s in nanocomposite, which corresponds to N-C-N, C-N and C-C bonds [70]. Then N 1s displayed two peaks at 399 and 397 eV, providing evidence for the presence of N 1s species (Figure S1c) with different environments of N-C and C-N-C bonds [71]. Cu 2p exposed two major peaks at 953 and 933 eV (Figure S1d) combined with two satellite lines, confirming the presence of Cu 2p species in the CN5CT composite [72]. The two peaks (Figure S1e) at 462 and 456 eV confirm the presence of Ti 2p species in CN5CT nanocomposite [72]. Finally, for O1s species, two strong peaks (Figure S1f) at 529 and 531 eV confirmed the existence of oxygen-metal and oxygen-hydrogen bonds in the composite.
The typical morphology of the g-C3N5 and g-C3N5/Cu3TiO4 nanocomposite is shown in Figure 3a,b, indicating that the prepared g-C3N5 has a sheet-like morphology. As seen in Figure 3a, exfoliation caused the evolution of 3-amino-1,2,4-triazole into a typical layered structure, showing the formation of a loosely packed 2D g-C3N5 structure. In Figure 3b, 10%CN5CT nanocomposite, and Cu3TiO4 nanoparticles are marked. The 10% CN5CT nanocomposites showed a strong connection between g-C3N5 and Cu3TiO4, indicating that there would be a physical adsorption. It also implied that g-C3N5 might prevent the aggregation of Cu3TiO4 NPs as compared to Figure 3b. The effectiveness of the electron-hole separation can be increased by closing interfacial contacts between g-C3N4 and Cu3TiO4 nanoparticles by excellent dispersion. According to the elemental mapping in Figure 3c–f, the elements C, N, Ti, Cu, and O were evenly distributed throughout the 2D layered and 3D nanoparticle structures.
The perfect morphology and structural nature of the g-C3N5 and 10% CN5CT were recorded by using TEM (Figure 4a,b). The agglomerated sheet-like structure of g-C3N5 was confirmed by a TEM image at 50 nm (Figure 4a). The surface of the g-C3N5 was smooth and had high porosity, which was confirmed by the sheet-like structure of the g-C3N5. The 3-AT was thermal disintegration, copolymerization, and elimination of CO2 and NH3, which may result in the development of g-C3N5 nanosheets, which have a smaller and thinner nanosheet like structure than g-C3N4. Both g-C3N4 and g-C3N5 are similar in appearance. The EDS of g-C3N5 provides information about the given sample, g-C3N5, which has a higher quantity of nitrogen atoms compared to carbon atoms (Figure 4c). On introducing Cu3TiO4 into g-C3N5 nanosheets, no changes were observed on the surface of g-C3N5. The Cu3TiO4 NPs are uniformly distributed on the g-C3N5 surface and form an effective heterojunction (Figure 4b). In Figure 4d, EDS analysis of 10%CNCT confirms the elements with weight percentage viz., N (52.4%), C (35.1%), O (8.7%), Ti (2%), and Cu (1.8%). Figure 4e depicts the SEAD pattern of the 10%CN5CT, demonstrating that the prepared composite has a good crystalline nature and corresponding planes of (111), (103), (102), and (106) in XRD. The average particle size of Cu3TiO4 (~26 nm) in 10%CN5CT was calculated and given in Figure 4f. From the TEM image, the surface area of the compound increased with the decrease in Cu3TiO4 particle size. The photocatalyst 10% CN5CT was found to coexist on the surface of g-C3N5 sheets. The Cu3TiO4 (dark-coloured) nanoparticles are uniformly scattered on the g-C3N5 matrix to produce an efficient heterojunction structure, with the average particle size of Cu3TiO4 being found to be around 50 nm. The large surface area enhances the photocatalytic activity. Due to their excellent surface area, both g-C3N5 and Cu3TiO4 have close contact with each other. Due to this nature, the 10%CN5CT nanocomposite has better photocatalytic activity.
The charge separation and transfer of the electron-hole pair were analysed by using photoluminescence (PL) emission spectroscopy. Figure 5a shows the PL results for the prepared CN, CT, and CN5CT at various loading percentages. The g-C3N5 nanosheets emission peak was observed at around ~470 nm with high intensity, which confirmed the high recombination rate. On the other hand, the introduction of Cu3TiO4 nanoparticles on the surface of g-C3N5 nanosheets reduced the intensity of the peak, which confirmed the suppression of recombination rate of the CN5CT nanocomposite. Notably, when compared to other CN5CT nanocomposite, the 10% CN5CT nanocomposite has a lower emission peak. The PL investigation is strong evidence that 10%CN5CT has less electron-hole pair recombination than 5%, 20%, and 30% of CN5CT.
The thermal stability of the prepared nanocomposite was carried out by using thermo gravimetric analysis (TGA). Here, in Figure 5b, is depicted the high stability of Cu3TiO4 nanoparticles, because no characteristic decomposition occurred in the Cu3TiO4 nanoparticles and up to 98% of Cu3TiO4 nanoparticles were getting residue at 800 oC, showing the high stability nature of nanoparticles. At 530 oC, g-C3N5 nanosheets began decomposing, and all of the carbon and nitrogen in g-C3N5 was easily eliminated above 700 °C (1% residue at 800 °C). In 10% CN5CT nanocomposite, g-C3N5 nanosheets got decomposed at 530 oC but the residue obtained was 12%, which shows the presence of Cu3TiO4 nanoparticles in g-C3N5 nanosheets at a 10% ratio was confirmed. The 10%CN5CT nanocomposite has improved the thermal stability compared to g-C3N5.

2.2. Photocatalytic Activity of g-C3N5/Cu3TiO4 Nanocomposite

In this study, we concentrated on the synthesis of different carbonitriles using a g-C3N5/Cu3TiO4 (CN5CT) photocatalyst. The catalyst exhibited better efficiency for the synthesis of carbonitriles in the presence of light. Ethanol has been used as solvent, which generated a yield of 94% (Table 1, entry 6). The reaction was carried out in the presence of BLUE LED visible light using a photocatalyst employing diketone, malanonitrile, and aldehyde (Scheme 1). Based on the characterization results, we investigated the CN5CT nanocomposite, which has a strong light absorption capacity and thereby increases the life time of electrons in the conduction band. The high electron-hole recombination rate improves the activity of the catalyst in the production of 1.4a. In this case, 20 mg of the catalyst was combined with the reactants and exposed to Blue LED light. Simultaneously, radicals formed on the catalyst surface, resulting in the formation of activated complexes. These are highly active species that are rapidly degraded to create products on the catalyst surface. In Table 1, the reaction conditions were adjusted to determine the optimal conditions for carbonitrile synthesis. Here, the solvent, catalyst, condition, and time were varied to determine the optimization condition. (Table 1 entries 1 and 3), produced yields (62% and 33%) that were significantly lower than the CN5CT nanocomposite. CN5CT nanocomposite and ethanol solvent worked better in all situations in the presence of light for 3 min. In all situations, the CN5CT nanocomposite and ethanol solvent worked better in all situations in the presence of light for 3 min. The increase in catalyst percentage also did not affect the product yield and for 5%, 20% and 30%, CN5CT nanocomposites give good yields (Table 1, entry 5, 7, 8).
The optimal condition from Table 1 was utilized to synthesize different carbonitrile derivatives, which is shown in Table 2. The physicochemical properties and evidence (FT-IR, 1H, 13C NMR, and GC-MS) of the synthesized carbonitrile derivatives 1.4a-g are shown in Supporting Information Figures S2–S29. Light excites the electrons on the catalyst surface, causing them to move into the excited conduction band. The radicals (O2 and OH) then react with the reactants to generate intermediates, which are then broken down to form the desired products on the catalyst surface. The catalyst, CN5CT nanocomposite, showed exceptional activity in the preparation of compound 1.4a. The recovered catalyst also performed admirably in the subsequent runs, with better yields.
In Table 2, we studied how several heterocycles were synthesized. We successfully prepared several substituted aldehydes such as electron donating, electron withdrawing, and halogenated aldehydes here. When donating substitutions were compared to others, they produced moderate yields (Table 2, compound 1.4b). In the proposed procedure, napthaldehyde and benzaldehyde substitutions produced the highest yields of all. The multi-component technique made good yields and cut down on the amount of time needed to make 1.4a.
The reaction was carried out using CN5CT and irradiated for 3 min in an open environment at RT with a BLUE LED light. Remarkably, 94% of the targeted product (1.4a) was achieved. Nevertheless, when ambient O2 was removed from the reaction mixture, the conversion was minimal and only a trace amount of product could be found if the reaction was completed in darkness or without photocatalysts. As a result, CN5CT photocatalyst, atmospheric oxygen, and BLUE LED visible light are all required for this transition.

2.3. Proposed Mechanism of 1.4a Synthesis

Based on the findings, a plausible mechanism was illustrated in Scheme 2. The more stable triplet state CN5CT* of the photoredox catalyst is then activated by the absorption of visible light and performs a single electron transfer (SET). This single electron transfer from 1.1 and 1.3 to the excited state of the CN5CT photocatalyst then lost a hydrogen atom and generated malononitrile radical (1.1a) and 1,3 cyclohexanedione radical (1.3a). The Knoevenagel condensation process between aldehyde (1.2) and activated malononitrile (1.1a) resulted in the synthesis of alkene intermediate (1.2a) [74]. Then, the intermediates 1.2b and 1.3a react in situ via Michael addition to give an adduct 1.3b, followed by intramolecular cyclisation of 1.3c, giving the desired product 1.4.

2.4. Transport Process of Photo-Excited Charge Carriers

The prepared samples’ light absorption in the 200–800 nm region was examined in order to obtain an appearance in the optical harvest property. The absorption edge of g-C3N5 is located at approximately 680 nm, as shown in Figure 6a. In contrast, the 10% CN5CT nanocomposite clearly showed a shift, which may have been caused by the intense interface interaction between Cu3TiO4 and g-C3N5. As a result, the representative 10 weight percentage (10%) CN5CT nanocomposite photocatalyst exhibits a greater capacity to capture visible light than 2D g-C3N5.
As seen in Figure 6b, the band-gap values were also calculated from the intercept of tangents to plots of (Ahv)1/2 versus photon energy. The optical band gap of g-C3N5, Cu3TiO4 and 10% CN5CT was calculated and found to be 2.02, 2.15, and 2.04 eV respectively. In comparison to pure g-C3N5 and Cu3TiO4, the 10% CN5CT nanocomposite exhibits a narrower band gap. According to the findings, the photocatalyst for the 10% CN5CT heterojunction complex has good reactivity to visible light, which may increase its photocatalytic activity. The results show that photo-generated electron-hole charge carriers are enhanced, which will boost photocatalytic activity compared to CN5 and CT.
To obtain the band levels of the g-C3N5/Cu3TiO4, the valance band (CB) and conduction band (CB) potentials for g-C3N5 & Cu3TiO4 were calculated with the aid of Equations (1) and (2) [75].
EVB = χ − Ee + ((1/2) * Eg)
ECB = EVB − Eg
where, Eg-band gap calculated from Figure 6b (2.02 eV for g-C3N5, and 2.15 eV for Cu3TiO4); ECB-CB potential; EVB-VB potential; Ee-energy of free electrons evaluated on the H scale (4.5 eV); χ–electro negativity (6.89 eV for g-C3N5 and 5.63 for Cu3TiO4) calculated from literature [64]. The calculated EVB and ECB were +1.38 and −0.64 eV for g-C3N5, +0.055 and −2.095 eV for Cu3TiO4, respectively.
It is well known that the heterojunction photocatalyst g-C3N5/Cu3TiO4 will be produced when g-C3N5 and Cu3TiO4 are coupled. According to the band gap structures of g-C3N5 & Cu3TiO4, the separation processes of photo-induced electron–hole (e/h+) is conveyed in Figure 7a,b, respectively.
According to the type-II heterojunction (Figure 7a), the holes of g-C3N5 formed by visible-light will transfer to the VB of Cu3TiO4 (0.055 eV). Altogether, es instantaneously migrate from the CB of Cu3TiO4 (−2.09 eV) to the CB of g-C3N5, (−0.64 eV). However, the potential energy (V vs. NHE) of es from g-C3N5 is much lower than O2/O2 (−0.33 V) [76], which designates that these electrons are unable to reducing O2 to ·O2. Correspondingly, holes of Cu3TiO4 have inadequate potential energy to oxidize H2O to ·OH (·OH/H2O = 2.34 V vs. NHE) [77]. Consequently, the type-II mechanism does not well define the photocatalytic activity by the g-C3N5/Cu3TiO4 heterojunction.
Our photocatalyst works in total tandem with the visible light wavelength from BLUE LED light sources. Both g-C3N5 and Cu3TiO4 can be stimulated to produce electrons and holes when exposed to light. The electrons in CB of g-C3N5 tend to migrate to VB of Cu3TiO4 and recombine with the holes there under the influence of the electron transport system, following the Z scheme charge transfer path. As a result, the highly redox active photogenerated electrons in CB of Cu3TiO4 and holes in VB of g-C3N5 are retained to participate in the photocatalytic activity. If the photo-induced charge transfer process follows the direct Z-Scheme mechanism as illustrated in Figure 7b, the photon-induced excited electrons would thermodynamically transfer from the CB of g-C3N5 to VB of Cu3TiO4. Meanwhile, the O2 could be reduced into ⋅O2 (reduction reaction) by the electrons generated by CB of Cu3TiO4 because of the more negative potential (−2.09 eV). Further, the holes of g-C3N5 are arrested by OH to produce the ˙OH, and the ˙OH are moved to the superficial of the nanocomposite to oxidize molecules, because of the more positive potential (+1.38 eV). This charge transfer process in the direct Z-scheme greatly improves the efficiency of photo-generated electron-hole separation, resulting in the enhancement of the photocatalytic activity [78,79,80].
The nanocomposite exhibited excellent stability, visible absorption nature, high yields, a shortened reaction time, and reusability (Figure S30a). The photocatalyst 10%CN5CT does not affected by the reaction mixture and giving higher yields for the next few runs. Then after recycled catalyst was showing good crystalline nature in XRD analysis shown in Figure S30b. In the presence of visible light, the narrow bandgap feature demonstrates a visible active photocatalyst. In visible light, green synthesis of different heterocycles using CN5CT nanocomposite produced good yields in a short period. The photocatalyst can be used repeatedly without affecting the activity of the catalyst.

3. Materials and Methods

3.1. Materials Information

All of the chemicals and reagents were obtained from Avra Chemicals Private Limited, which is located in Andhra Pradesh, India. The column solvents and other chemicals were brought from Avra chemicals and used without further purification or processing. We purchased the lemons from the Vellore local market in Katpadi, Tamil Nadu, India. The NMR solvents came from Merck and was said to be 99.9% pure by the manufacturer.

3.2. Preparation of Citrus Limon Extract

Fresh lemons were split into half and squeezed to gather the pulp. There is some excess waste in the gathered juice (seeds and peel). To eliminate the additional trash, the pulp was centrifuged at 10,000 rpm. The clear liquid from the top layer was collected and it was stored in refrigerator.

3.3. Preparation of Cu3TiO4 NPs

Copper acetate (1 M) and titanium tert-butoxide (1 M) were taken in two separate beakers with 50mL of deionized water. A titanium tert-butoxide solution was added into the copper acetate solution with constant stirring. Then 5 mL of citrus extract obtained was added and the mixture was kept stirring overnight. Bluish-white coloured residue was separated and washed with a water-ethanol solution to remove the other impurities. The precipitate was heated at 500 °C for 4 h in a silica crucible. Finally, a black precipitate was obtained and subjected to characterization.

3.4. Preparation of g-C3N5 Nanosheets and Composite

In a beaker, 1 mmol of 3-amino-1,2,4-triazole was stirred well with 20 pellets of NaOH for 20 min. The colour of the mixture was altered to light brown after it was ground. For 4 h, the mixture was heated at 550 °C at a rate of 20 °C/min. Finally, grey g-C3N5 was collected and rinsed 10 times with distilled water to eliminate unreacted NaOH before being kept for characterization. The 5%, 10%, 20%, and 30% CN5CT were produced after processing g-C3N5 with Cu3TiO4 at various concentrations.

3.5. General Synthesis of Carbonitriles (1.4a-g) under Visible Light

1,3-cyclohexadione (1 mmol), corresponding aldehydes (1 mmol), malononitrile (1 mmol), CN5CT photocatalyst (20 mg), and 5 mL of ethanol were placed into a 25 mL round-bottom flask in the BLUE LED (12 W, 450 to 495 wavelength) light chamber. The reaction mixture was illuminated, and TLC was tested every minute to identify the result. After reaction completion, the reaction mixture was poured into crushed ice and precipitated out. The solids were removed from the ethanol and recrystallized.

4. Conclusions

In conclusion, we described the fabrication of direct Z-scheme Cu3TiO4/g-C3N5 heterojunction and their characterization. The Cu3TiO4/g-C3N5 photocatalyst has expressed excellent activity for the synthesis of carbonitriles derivatives. In an as-prepared heterojunction, 10% Cu3TiO4 is dispersed on surface the surface of g-C3N5 exhibited better photocatalytic activity. When compared to pure Cu3TiO4 and g-C3N5, Cu3TiO4/g-C3N5 nanocomposites perform better at photocatalysis due to their distinctive nanostructure and direct Z-scheme heterojunction. Low-energy consumption visible light BLUE LED lights were used for the irradiation. The reaction has been carried out immediately and the photocatalyst was showing excellent activity, reusability, easily recoverable, less recombination rate in visible light and high surface area.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal12121593/s1, Figure S1: XPS of Cu3TiO4/g-C3N5 nanocomposites; Figures S2–S29: FT-IR, 1H, 13C NMR and GC-MS of carbonitrile derivatives 1.4a–g; Figure S30: Reusability and reused XRD.

Author Contributions

Original draft preparation, data collection and writing, M.A. and T.C.; Writing and editing, R.M.; Data analysis, K.A.; Writing-review and editing, G.M. and S.M.R.; Supervision, S.M.R. All authors discussed the results and contributed to the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

Our profound thanks go to the management of the Vellore Institute of Technology for providing the Sophisticated Instrument Facility (SIF) to aid in the research endeavour.

Conflicts of Interest

The authors declared no potential conflicts of interest.

References

  1. Nazeri, M.T.; Shaabani, A. Synthesis of Polysubstituted Pyrroles via Isocyanide-Based Multicomponent Reactions as an Efficient Synthesis Tool. New J. Chem. 2021, 45, 21967–22011. [Google Scholar] [CrossRef]
  2. Chopra, P.K.P.G.; Lambat, T.L.; Mahmood, S.H.; Chaudhary, R.G.; Banerjee, S. Sulfamic Acid as Versatile Green Catalyst Used for Synthetic Organic Chemistry: A Comprehensive Update. Chem. Sel. 2021, 6, 6867–6889. [Google Scholar] [CrossRef]
  3. Dekamin, M.G.; Eslami, M. Highly Efficient Organocatalytic Synthesis of Diverse and Densely Functionalized 2-Amino-3-Cyano-4H-Pyrans under Mechanochemical Ball Milling. Green Chem. 2014, 16, 4914–4921. [Google Scholar] [CrossRef]
  4. Wang, X.; Yin, F.; Bi, Y.; Cheng, G.; Li, J.; Hou, L.; Li, Y.; Yang, B.; Liu, W.; Yang, L. Rapid and Sensitive Detection of Zika Virus by Reverse Transcription Loop-Mediated Isothermal Amplification. J. Virol. Methods 2016, 238, 86–93. [Google Scholar] [CrossRef]
  5. Singh, S.B. Copper Nanocatalysis in Multi-Component Reactions: A Green to Greener Approach. Curr. Catal. 2018, 7, 80–88. [Google Scholar] [CrossRef]
  6. Ghamari Kargar, P.; Bagherzade, G. Robust, Highly Active, and Stable Supported Co(ii) Nanoparticles on Magnetic Cellulose Nanofiber-Functionalized for the Multi-Component Reactions of Piperidines and Alcohol Oxidation. RSC Adv. 2021, 11, 23192–23206. [Google Scholar] [CrossRef]
  7. Kamanna, K.; Khatavi, S.Y. Microwave-Accelerated Carbon-Carbon and Carbon-Heteroatom Bond Formation via Multi-Component Reactions: A Brief Overview. Curr. Microw. Chem. 2020, 7, 23–39. [Google Scholar] [CrossRef]
  8. Ramesh, R.; Maheswari, S.; Malecki, J.G.; Lalitha, A. NaN3 Catalyzed Highly Convenient Access to Functionalized 4 H -Chromenes: A Green One-Pot Approach for Diversity Amplification. Polycycl. Aromat. Compd. 2020, 40, 1581–1594. [Google Scholar] [CrossRef]
  9. Majumdar, N.; Paul, N.D.; Mandal, S.; de Bruin, B.; Wulff, W.D. Catalytic Synthesis of 2 H -Chromenes. ACS Catal. 2015, 5, 2329–2366. [Google Scholar] [CrossRef]
  10. Hauguel, C.; Ducellier, S.; Provot, O.; Ibrahim, N.; Lamaa, D.; Balcerowiak, C.; Letribot, B.; Nascimento, M.; Blanchard, V.; Askenatzis, L.; et al. Design, Synthesis and Biological Evaluation of Quinoline-2-Carbonitrile-Based Hydroxamic Acids as Dual Tubulin Polymerization and Histone Deacetylases Inhibitors. Eur. J. Med. Chem. 2022, 240, 114573. [Google Scholar] [CrossRef]
  11. Khan, S.A.; Asiri, A.M.; Basisi, H.M.; Asad, M.; Zayed, M.E.M.; Sharma, K.; Wani, M.Y. Synthesis and Evaluation of Quinoline-3-Carbonitrile Derivatives as Potential Antibacterial Agents. Bioorg. Chem. 2019, 88, 102968. [Google Scholar] [CrossRef] [PubMed]
  12. Patel, K.B.; Kumari, P. A Review: Structure-Activity Relationship and Antibacterial Activities of Quinoline Based Hybrids. J. Mol. Struct. 2022, 1268, 133634. [Google Scholar] [CrossRef]
  13. Liu, T.; Wu, Z.; He, Y.; Xiao, Y.; Xia, C. Single and Dual Target Inhibitors Based on Bcl-2: Promising Anti-Tumor Agents for Cancer Therapy. Eur. J. Med. Chem. 2020, 201, 112446. [Google Scholar] [CrossRef]
  14. Sharma, D.; Kumar, M.; Das, P. Application of Cyclohexane-1,3-Diones for Six-Membered Oxygen-Containing Heterocycles Synthesis. Bioorg. Chem. 2021, 107, 104559. [Google Scholar] [CrossRef] [PubMed]
  15. Oliveira-Pinto, S.; Pontes, O.; Baltazar, F.; Costa, M. In Vivo Efficacy Studies of Chromene-Based Compounds in Triple-Negative Breast Cancer—A Systematic Review. Eur. J. Pharmacol. 2020, 887, 173452. [Google Scholar] [CrossRef]
  16. Costa, M.; Dias, T.A.; Brito, A.; Proença, F. Biological Importance of Structurally Diversified Chromenes. Eur. J. Med. Chem. 2016, 123, 487–507. [Google Scholar] [CrossRef]
  17. Sadek, K.U.; Mekheimer, R.A.H.; Abd-Elmonem, M.; Abdel-Hameed, A.; Elnagdi, M.H. Recent Developments in the Enantioselective Synthesis of Polyfunctionalized Pyran and Chromene Derivatives. Tetrahedron Asymmetry 2017, 28, 1462–1485. [Google Scholar] [CrossRef]
  18. Chaudhary, A.; Singh, K.; Verma, N.; Kumar, S.; Kumar, D.; Sharma, P.P. Chromenes—A Novel Class of Heterocyclic Compounds: Recent Advancements and Future Directions. Mini.-Rev. Med. Chem. 2022, 22, 2736–2751. [Google Scholar] [CrossRef]
  19. Son, Y.-A.; Gwon, S.-Y.; Kim, S.-H. Chromene and Imidazole Based D-π-A Chemosensor Preparation and Its Anion Responsive Effects. Mol. Cryst. Liq. Cryst. 2014, 599, 16–22. [Google Scholar] [CrossRef]
  20. Dagilienė, M.; Markuckaitė, G.; Krikštolaitytė, S.; Šačkus, A.; Martynaitis, V. Cyanide Anion Determination Based on Nucleophilic Addition to 6-[(E)-(4-Nitrophenyl)Diazenyl]-1′,3,3′,4-Tetrahydrospiro[Chromene-2,2′-Indole] Derivatives. Chem. Sens. 2022, 10, 185. [Google Scholar] [CrossRef]
  21. Kar, S.; Sanderson, H.; Roy, K.; Benfenati, E.; Leszczynski, J. Green Chemistry in the Synthesis of Pharmaceuticals. Chem. Rev. 2022, 122, 3637–3710. [Google Scholar] [CrossRef] [PubMed]
  22. Clarke, C.J.; Tu, W.-C.; Levers, O.; Bröhl, A.; Hallett, J.P. Green and Sustainable Solvents in Chemical Processes. Chem. Rev. 2018, 118, 747–800. [Google Scholar] [CrossRef]
  23. Kamanna, K.; Amaregouda, Y. Synthesis of Bioactive Scaffolds Catalyzed by Agro-Waste-Based Solvent Medium. Phys. Sci. Rev. 2022. [Google Scholar] [CrossRef]
  24. Védrine, J. Heterogeneous Catalysis on Metal Oxides. Catalysts 2017, 7, 341. [Google Scholar] [CrossRef] [Green Version]
  25. Chellapandi, T.; Madhumitha, G. Montmorillonite Clay-Based Heterogenous Catalyst for the Synthesis of Nitrogen Heterocycle Organic Moieties: A Review. Mol. Divers. 2021. [CrossRef] [PubMed]
  26. Singh, B.K.; Lee, S.; Na, K. An Overview on Metal-Related Catalysts: Metal Oxides, Nanoporous Metals and Supported Metal Nanoparticles on Metal Organic Frameworks and Zeolites. Rare Met. 2020, 39, 751–766. [Google Scholar] [CrossRef]
  27. Shifrina, Z.B.; Matveeva, V.G.; Bronstein, L.M. Role of Polymer Structures in Catalysis by Transition Metal and Metal Oxide Nanoparticle Composites. Chem. Rev. 2020, 120, 1350–1396. [Google Scholar] [CrossRef]
  28. Terna, A.D.; Elemike, E.E.; Mbonu, J.I.; Osafile, O.E.; Ezeani, R.O. The Future of Semiconductors Nanoparticles: Synthesis, Properties and Applications. Mater. Sci. Eng. B 2021, 272, 115363. [Google Scholar] [CrossRef]
  29. Ge, M.Z.; Cao, C.Y.; Li, S.H.; Tang, Y.X.; Wang, L.N.; Qi, N.; Lai, Y.K. In situ plasmonic Ag nanoparticle anchored TiO2 nanotube arrays as visible-light-driven photocatalysts for enhanced water splitting. Nanoscale 2016, 8, 5226–5234. [Google Scholar] [CrossRef]
  30. Ge, M.; Cai, J.; Iocozzia, J.; Cao, C.; Huang, J.; Zhang, X.; Lin, Z. A review of TiO2 nanostructured catalysts for sustainable H2 generation. Int. J. Hydrogren Energy 2017, 42, 8418–8449. [Google Scholar] [CrossRef]
  31. Ge, M.; Li, Q.; Cao, C.; Huang, J.; Li, S.; Zhang, S.; Lai, Y. One-dimensional TiO2 nanotube photocatalysts for solar water splitting. Adv. Sci. 2017, 4, 1600152. [Google Scholar] [CrossRef] [PubMed]
  32. Hu, X.; Luo, T.; Lin, Y.; Yang, M. Construction of Novel Z-Scheme g-C3N4/AgBr-Ag Composite for Efficient Photocatalytic Degradation of Organic Pollutants under Visible Light. Catalysts 2022, 12, 1309. [Google Scholar] [CrossRef]
  33. Cuong, H.N.; Pansambal, S.; Ghotekar, S.; Oza, R.; Thanh Hai, N.T.; Viet, N.M.; Nguyen, V.-H. New Frontiers in the Plant Extract Mediated Biosynthesis of Copper Oxide (CuO) Nanoparticles and Their Potential Applications: A Review. Environ. Res. 2022, 203, 111858. [Google Scholar] [CrossRef] [PubMed]
  34. Kumar, A.; Choudhary, P.; Kumar, A.; Camargo, P.H.C.; Krishnan, V. Recent Advances in Plasmonic Photocatalysis Based on TiO2 and Noble Metal Nanoparticles for Energy Conversion, Environmental Remediation, and Organic Synthesis. Small 2022, 18, 2101638. [Google Scholar] [CrossRef]
  35. Gnanasekaran, L.; Pachaiappan, R.; Kumar, P.S.; Hoang, T.K.A.; Rajendran, S.; Durgalakshmi, D.; Soto-Moscoso, M.; Cornejo-Ponce, L.; Gracia, F. Visible Light Driven Exotic p (CuO)–n (TiO2) Heterojunction for the Photodegradation of 4-Chlorophenol and Antibacterial Activity. Environ. Pollut. 2021, 287, 117304. [Google Scholar] [CrossRef]
  36. Marzo, L.; Pagire, S.K.; Reiser, O.; König, B. Visible-Light Photocatalysis: Does It Make a Difference in Organic Synthesis? Angew. Chem. Int. Ed. 2018, 57, 10034–10072. [Google Scholar] [CrossRef]
  37. Alaghmandfard, A.; Ghandi, K. A Comprehensive Review of Graphitic Carbon Nitride (g-C3N4)–Metal Oxide-Based Nanocomposites: Potential for Photocatalysis and Sensing. Nanomaterials 2022, 12, 294. [Google Scholar] [CrossRef]
  38. Arunachalapandi, M.; Roopan, S.M. Environment Friendly G-C3N4-Based Catalysts and Their Recent Strategy in Organic Transformations. High Energy Chem. 2022, 56, 73–90. [Google Scholar] [CrossRef]
  39. Mun, S.J.; Park, S.-J. Graphitic Carbon Nitride Materials for Photocatalytic Hydrogen Production via Water Splitting: A Short Review. Catalysts 2019, 9, 805. [Google Scholar] [CrossRef] [Green Version]
  40. Chen, J.; Fang, S.; Shen, Q.; Fan, J.; Li, Q.; Lv, K. Recent Advances of Doping and Surface Modifying Carbon Nitride with Characterization Techniques. Catalysts 2022, 12, 962. [Google Scholar] [CrossRef]
  41. Wang, J.; Wang, S. A Critical Review on Graphitic Carbon Nitride (g-C3N4)-Based Materials: Preparation, Modification and Environmental Application. Coord. Chem. Rev. 2022, 453, 214338. [Google Scholar] [CrossRef]
  42. Mane, G.P.; Talapaneni, S.N.; Lakhi, K.S.; Ilbeygi, H.; Ravon, U.; Al-Bahily, K.; Mori, T.; Park, D.H.; Vinu, A. Highly ordered nitrogen-rich mesoporous carbon nitrides and their superior performance for sensing and photocatalytic hydrogen generation. Angew. Chem. Int. Ed. 2017, 56, 8481–8485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Liu, D.; Yao, J.; Chen, S.; Zhang, J.; Li, R.; Peng, T. Construction of rGO-coupled C3N4/C3N5 2D/2D Z-scheme heterojunction to accelerate charge separation for efficient visible light H2 evolution. Appl. Catal. B. 2022, 318, 121822. [Google Scholar] [CrossRef]
  44. Luo, T.; Hu, X.; She, Z.; Wei, J.; Feng, X.; Chang, F. Synergistic effects of Ag-doped and morphology regulation of graphitic carbon nitride nanosheets for enhanced photocatalytic performance. J. Mol. Liq. 2021, 324, 114772. [Google Scholar] [CrossRef]
  45. Li, S.; Cai, M.; Liu, Y.; Wang, C.; Yan, R.; Chen, X. Constructing Cd0.5Zn0.5S/Bi2WO6 S-scheme heterojunction for boosted photocatalytic antibiotic oxidation and Cr (VI) reduction. Adv. Powder Technol. 2023, 2, 100073. [Google Scholar] [CrossRef]
  46. Li, Z.; Li, H.; Wang, S.; Yang, F.; Zhou, W. Mesoporous black TiO2/MoS2/Cu2S hierarchical tandem heterojunctions toward optimized photothermal-photocatalytic fuel production. J. Chem. Eng. 2022, 427, 131830. [Google Scholar] [CrossRef]
  47. Wang, S.; Guan, B.Y.; Lou, X.W.D. Construction of ZnIn2S4–In2O3 hierarchical tubular heterostructures for efficient CO2 photoreduction. J. Am. Chem. Soc. 2018, 140, 5037–5040. [Google Scholar] [CrossRef]
  48. Xiao, Y.; Guo, S.; Tian, G.; Jiang, B.; Ren, Z.; Tian, C.; Li, W.; Fu, H. Synergetic enhancement of surface reactions and charge separation over holey C3N4/TiO2 2D heterojunctions. Sci. Bull. 2021, 66, 275–283. [Google Scholar] [CrossRef]
  49. Humayun, M.; Ullah, H.; Tahir, A.A.; bin Mohd Yusoff, A.R.; Mat Teridi, M.A.; Nazeeruddin, M.K.; Luo, W. An Overview of the Recent Progress in Polymeric Carbon Nitride Based Photocatalysis. Chem. Rec. 2021, 21, 1811–1844. [Google Scholar] [CrossRef]
  50. Palanivel, B.; Mani, A. Conversion of a Type-II to a Z-Scheme Heterojunction by Intercalation of a 0D Electron Mediator between the Integrative NiFe2O4/g-C3N4 Composite Nanoparticles: Boosting the Radical Production for Photo-Fenton Degradation. ACS Omega 2020, 5, 19747–19759. [Google Scholar] [CrossRef]
  51. Lu, L.; Wang, G.; Zou, M.; Wang, J.; Li, J. Effects of Calcining Temperature on Formation of Hierarchical TiO2/g-C3N4 Hybrids as an Effective Z-Scheme Heterojunction Photocatalyst. Appl. Surf. Sci. 2018, 441, 1012–1023. [Google Scholar] [CrossRef]
  52. Wang, M.; Jin, C.; Kang, J.; Liu, J.; Tang, Y.; Li, Z.; Li, S. CuO/g-C3N4 2D/2D Heterojunction Photocatalysts as Efficient Peroxymonosulfate Activators under Visible Light for Oxytetracycline Degradation: Characterization, Efficiency and Mechanism. Chem. Eng. J. 2021, 416, 128118. [Google Scholar] [CrossRef]
  53. Shen, D.; Li, X.; Ma, C.; Zhou, Y.; Sun, L.; Yin, S.; Huo, P.; Wang, H. Synthesized Z-Scheme Photocatalyst ZnO/g-C3N4 for Enhanced Photocatalytic Reduction of CO2. New J. Chem. 2020, 44, 16390–16399. [Google Scholar] [CrossRef]
  54. Meng, J.; Wang, X.; Liu, Y.; Ren, M.; Zhang, X.; Ding, X.; Guo, Y.; Yang, Y. Acid-Induced Molecule Self-Assembly Synthesis of Z-Scheme WO3/g-C3N4 Heterojunctions for Robust Photocatalysis against Phenolic Pollutants. Chem. Eng. J. 2021, 403, 126354. [Google Scholar] [CrossRef]
  55. Dadigala, R.; Bandi, R.; Gangapuram, B.R.; Dasari, A.; Belay, H.H.; Guttena, V. Fabrication of Novel 1D/2D V2O5/g-C3N4 Composites as Z-Scheme Photocatalysts for CR Degradation and Cr (VI) Reduction under Sunlight Irradiation. J. Environ. Chem. Eng. 2019, 7, 102822. [Google Scholar] [CrossRef]
  56. Wu, B.; Li, Y.; Su, K.; Tan, L.; Liu, X.; Cui, Z.; Yang, X.; Liang, Y.; Li, Z.; Zhu, S.; et al. The Enhanced Photocatalytic Properties of MnO2/g-C3N4 Heterostructure for Rapid Sterilization under Visible Light. J. Hazard. Mater. 2019, 377, 227–236. [Google Scholar] [CrossRef]
  57. He, R.; Zhou, J.; Fu, H.; Zhang, S.; Jiang, C. Room-Temperature in Situ Fabrication of Bi2O3/g-C3N4 Direct Z-Scheme Photocatalyst with Enhanced Photocatalytic Activity. Appl. Surf. Sci. 2018, 430, 273–282. [Google Scholar] [CrossRef]
  58. Yin, H.; Cao, Y.; Fan, T.; Zhang, M.; Yao, J.; Li, P.; Chen, S.; Liu, X. In Situ Synthesis of Ag3PO4/C3N5 Z-Scheme Heterojunctions with Enhanced Visible-Light-Responsive Photocatalytic Performance for Antibiotics Removal. Sci. Total Environ. 2021, 754, 141926. [Google Scholar] [CrossRef]
  59. Vadivel, S.; Fujii, M.; Rajendran, S. Facile Synthesis of Broom Stick like FeOCl/g-C3N5 Nanocomposite as Novel Z-Scheme Photocatalysts for Rapid Degradation of Pollutants. Chemosphere 2022, 307, 135716. [Google Scholar] [CrossRef]
  60. Chu, W.; Li, S.; Zhou, H.; Shi, M.; Zhu, J.; He, P.; Liu, W.; Wu, M.; Wu, J.; Yan, X. A Novel Bionic Flower-like Z-Scheme Bi4O5I2/g-C3N5 Heterojunction with I/I3 active sites and π-conjugated structure for increasing photocatalytic oxidation of elemental mercury. J. Environ. Chem. Eng. 2022, 10, 108623. [Google Scholar] [CrossRef]
  61. Wang, R.; Zhang, K.; Zhong, X.; Jiang, F. Z-Scheme LaCoO3/C3N5 for Efficient Full-Spectrum Light-Simulated Solar Photocatalytic Hydrogen Generation. RSC Adv. 2022, 12, 24026–24036. [Google Scholar] [CrossRef]
  62. Rajendran, S.; Chellapandi, T.; UshaVipinachandran, V.; Venkata Ramanaiah, D.; Dalal, C.; Sonkar, S.K.; Madhumitha, G.; Bhunia, S.K. Sustainable 2D Bi2WO6/g-C3N5 Heterostructure as Visible Light-Triggered Abatement of Colorless Endocrine Disruptors in Wastewater. Appl. Surf. Sci. 2022, 577, 151809. [Google Scholar] [CrossRef]
  63. Deng, S.; Li, Z.; Zhao, T.; Huang, G.; Wang, J.; Bi, J. Direct Z-Scheme Covalent Triazine-Based Framework/Bi2WO6 Heterostructure for Efficient Photocatalytic Degradation of Tetracycline: Kinetics, Mechanism and Toxicity. J. Water Process Eng. 2022, 49, 103021. [Google Scholar] [CrossRef]
  64. Das, A. LED Light Sources in Organic Synthesis: An Entry to a Novel Approach. Lett. Org. Chem. 2022, 19, 283–292. [Google Scholar] [CrossRef]
  65. Cai, Z.; Huang, Y.; Ji, H.; Liu, W.; Fu, J.; Sun, X. Type-II Surface Heterojunction of Bismuth-Rich Bi4O5Br2 on Nitrogen-Rich g-C3N5 Nanosheets for Efficient Photocatalytic Degradation of Antibiotics. Sep. Purif. Technol. 2022, 280, 119772. [Google Scholar] [CrossRef]
  66. Arunachalapandi, M.; Roopan, S.M. Ultrasound/Visible Light-Mediated Synthesis of N-Heterocycles Using g-C3N4/Cu3TiO4 as Sonophotocatalyst. Res. Chem. Intermed. 2021, 47, 3363–3378. [Google Scholar] [CrossRef]
  67. Wei, W.; Gong, H.; Sheng, L.; Wu, H.; Zhu, S.; Feng, L.; Li, X.; You, W. Highly Efficient Photocatalytic Activity and Mechanism of Novel Er3+ and Tb3+ Co-Doped BiOBr/g-C3N5 towards Sulfamethoxazole Degradation. Ceram. Int. 2021, 47, 24062–24072. [Google Scholar] [CrossRef]
  68. Vadivel, S.; Fujii, M.; Rajendran, S. Novel S-Scheme 2D/2D Bi4O5Br2 Nanoplatelets/g-C3N5 Heterojunctions with Enhanced Photocatalytic Activity towards Organic Pollutants Removal. Environ. Res. 2022, 213, 113736. [Google Scholar] [CrossRef]
  69. Ramezanalizadeh, H.; Rafiee, E. Design, Fabrication, Electro- and Photoelectrochemical Investigations of Novel CoTiO3/CuBi2O4 Heterojunction Semiconductor: An Efficient Photocatalyst for the Degradation of DR16 Dye. Mater. Sci. Semicond. Process. 2020, 113, 105055. [Google Scholar] [CrossRef]
  70. Li, M.; Lu, Q.; Liu, M.; Yin, P.; Wu, C.; Li, H.; Zhang, Y.; Yao, S. Photsoinduced Charge Separation via the Double-Electron Transfer Mechanism in Nitrogen Vacancies g-C3N5/BiOBr for the Photoelectrochemical Nitrogen Reduction. ACS Appl. Mater. Interfaces 2020, 12, 38266–38274. [Google Scholar] [CrossRef]
  71. Rajendran, R.; Vignesh, S.; Suganthi, S.; Raj, V.; Kavitha, G.; Palanivel, B.; Shkir, M.; Algarni, H. G-C3N4/TiO2/CuO S-Scheme Heterostructure Photocatalysts for Enhancing Organic Pollutant Degradation. J. Phys. Chem. Solids 2022, 161, 110391. [Google Scholar] [CrossRef]
  72. Saravanakumar, K.; Maheskumar, V.; Yea, Y.; Yoon, Y.; Muthuraj, V.; Park, C.M. 2D/2D nitrogen-rich graphitic carbon nitride coupled Bi2WO6 S-scheme heterojunction for boosting photodegradation of tetracycline: Influencing factors, intermediates, and insights into the mechanism. Compos. Part B Eng. 2022, 234, 109726. [Google Scholar] [CrossRef]
  73. Chen, L.; Bao, S.; Yang, L.; Zhang, X.; Li, B.; Li, Y. Cheap Thiamine Hydrochloride as Efficient Catalyst for Synthesis of 4H-Benzo[b]Pyrans in Aqueous Ethanol. Res. Chem. Intermed. 2017, 43, 3883–3891. [Google Scholar] [CrossRef]
  74. Tabassum, S.; Govindaraju, S.; Kendrekar, P. (Mes-Acr-Me)+ClO4– Catalyzed Visible Light-Supported, One-Pot Green Synthesis of 1,8-Naphthyridine-3-Carbonitriles. Top. Catal. 2022, 1–9. [Google Scholar] [CrossRef]
  75. Aditya, M.N.; Chellapandi, T.; Prasad, G.K.; Venkatesh, M.J.P.; Khan, M.M.R.; Madhumitha, G.; Roopan, S.M. Biosynthesis of Rod Shaped Gd2O3 on G-C3N4 as Nanocomposite for Visible Light Mediated Photocatalytic Degradation of Pollutants and RSM Optimization. Diam. Relat. Mater. 2022, 121, 108790. [Google Scholar] [CrossRef]
  76. Mondal, M.; Banerjee, S.; Mal, S.; Das, S.; Pradhan, S.K. Nanocomposites of GaBr3 and BiBr3 Nanocrystals on BiOBr for the Photocatalytic Degradation of Dyes and Tetracycline. ACS Appl. Nano. Mater. 2022, 5, 15676–15691. [Google Scholar] [CrossRef]
  77. Deng, S.; Yang, Z.; Lv, G.; Zhu, Y.; Li, H.; Wang, F.; Zhang, X. WO3 Nanosheets/g-C3N4 Nanosheets’ Nanocomposite as an Effective Photocatalyst for Degradation of Rhodamine B. Appl. Phys. A 2019, 125, 44. [Google Scholar] [CrossRef]
  78. Miao, Z.; Zhang, Y.; Wang, N.; Xu, P.; Wang, X. BiOBr/Bi2S3 Heterojunction with S-Scheme Structureand Oxygen Defects: In-Situ Construction and Photocatalytic Behavior for Reduction of CO2 with H2O. J. Colloid Interface Sci. 2022, 620, 407–418. [Google Scholar] [CrossRef]
  79. Xing, P.; Zhou, F.; Li, Z. Preparation of WO3/g-C3N4 Composites with Enhanced Photocatalytic Hydrogen Production Performance. Appl. Phys. A 2019, 125, 788. [Google Scholar] [CrossRef]
  80. Ye, R.; Fang, H.; Zheng, Y.-Z.; Li, N.; Wang, Y.; Tao, X. Fabrication of CoTiO3/g-C3N4 Hybrid Photocatalysts with Enhanced H2 Evolution: Z-Scheme Photocatalytic Mechanism Insight. ACS Appl. Mater. Interfaces 2016, 8, 13879–13889. [Google Scholar] [CrossRef]
Figure 1. XRD images of prepared CN, CT and various percentages CNCT nanocomposites.
Figure 1. XRD images of prepared CN, CT and various percentages CNCT nanocomposites.
Catalysts 12 01593 g001
Figure 2. (a) BET isotherm and (b) FT-IR spectra.
Figure 2. (a) BET isotherm and (b) FT-IR spectra.
Catalysts 12 01593 g002
Figure 3. FE-SEM images of (a) g-C3N5 (b) 10% CN5CT nanocomposite and (cg) elemental mapping of 10% CN5CT.
Figure 3. FE-SEM images of (a) g-C3N5 (b) 10% CN5CT nanocomposite and (cg) elemental mapping of 10% CN5CT.
Catalysts 12 01593 g003
Figure 4. (a,b) TEM images, (c,d) EDS spectra of g-C3N5, 10% CN5CT nanocomposite (e) SEAD pattern of 10% CN5CT. (f) Average particle size of Cu3TiO4 nanoparticles in 10% CN5CT.
Figure 4. (a,b) TEM images, (c,d) EDS spectra of g-C3N5, 10% CN5CT nanocomposite (e) SEAD pattern of 10% CN5CT. (f) Average particle size of Cu3TiO4 nanoparticles in 10% CN5CT.
Catalysts 12 01593 g004
Figure 5. (a) PL emission spectra of g-C3N5, 5% CN5CT, 30% CN5CT, 20% CN5CT, 10% CN5CT (b) TGA analysis of Cu3TiO4, g-C3N5, and 10% CN5CT nanocomposite.
Figure 5. (a) PL emission spectra of g-C3N5, 5% CN5CT, 30% CN5CT, 20% CN5CT, 10% CN5CT (b) TGA analysis of Cu3TiO4, g-C3N5, and 10% CN5CT nanocomposite.
Catalysts 12 01593 g005
Scheme 1. Synthesis of hexahydroquinoline-2-carbonitrile using CN5CT nanocomposite in Blue LED light.
Scheme 1. Synthesis of hexahydroquinoline-2-carbonitrile using CN5CT nanocomposite in Blue LED light.
Catalysts 12 01593 sch001
Scheme 2. Mechanism of hexahydroquinoline-2-carbonitrile using g-C3N5/Cu3TiO4.
Scheme 2. Mechanism of hexahydroquinoline-2-carbonitrile using g-C3N5/Cu3TiO4.
Catalysts 12 01593 sch002
Figure 6. (a) UV-Vis DRS spectra and (b) bandgap plot.
Figure 6. (a) UV-Vis DRS spectra and (b) bandgap plot.
Catalysts 12 01593 g006
Figure 7. Proposed electron-hole pair transfer mechanism of g-C3N5/Cu3TiO4 nanocomposite (a) type-II, (b) Direct Z-Scheme.
Figure 7. Proposed electron-hole pair transfer mechanism of g-C3N5/Cu3TiO4 nanocomposite (a) type-II, (b) Direct Z-Scheme.
Catalysts 12 01593 g007
Table 1. Optimization for the synthesis of Carbonitriles in BLUE LED visible light.
Table 1. Optimization for the synthesis of Carbonitriles in BLUE LED visible light.
S.NoCatalystSolventConditionsTime (minutes)Yield a (%)
1.CuOEtOHLight 6062
2.CuOAcEtOHLight 6066
3.TiO2EtOHLight 6033
4.Cu3TiO4EtOHLight 6075
5.5% CN5CTEtOHLight367
6.10% CN5CTEtOHLight3 94
7.20% CN5CTEtOHLight3 81
8.30% CN5CTEtOHLight389
9.10% CN5CTH2OLight1061
10.10% CN5CTTHFRT6023
11.10% CN5CTAcetonitrileRT6021
1210%CN5CTEtOHRT1095
13.Without CatalystEtOHRT10Trace
14.10% CN5CTSolvent freeHeat1078
Diketone (1 mmol), aldehyde (1 mmol), malononitrile (1 mmol), ethanol, photocatalyst (20 mg). a Yields—Isolated yields.
Table 2. Scope of different substrate.
Table 2. Scope of different substrate.
Catalysts 12 01593 i001NameRTimeTON aTOF aYield bRef. c
1.4a-Ph58653.5285[73]
1.4b4-(OCH3)-Ph58323.2682[73]
1.4c2-NO2-Ph38763.7888[73]
1.4d4-OH-Ph58313.2281[73]
1.4e4-F-Ph58783.8690[73]
1.4f2-Cl-Ph39234.6394[73]
1.4g2-Nap39024.1292[73]
Diketone (1 mmol), aldehyde (1 mmol), malononitrile (1 mmol), ethanol, photocatalyst (20 mg). a TON—Turn Over Number; TOF—Turn Over Frequency. b Yields—Isolated yields. c—reference.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Arunachalapandi, M.; Chellapandi, T.; Madhumitha, G.; Manjupriya, R.; Aravindraj, K.; Roopan, S.M. Direct Z-Scheme g-C3N5/Cu3TiO4 Heterojunction Enhanced Photocatalytic Performance of Chromene-3-Carbonitriles Synthesis under Visible Light Irradiation. Catalysts 2022, 12, 1593. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12121593

AMA Style

Arunachalapandi M, Chellapandi T, Madhumitha G, Manjupriya R, Aravindraj K, Roopan SM. Direct Z-Scheme g-C3N5/Cu3TiO4 Heterojunction Enhanced Photocatalytic Performance of Chromene-3-Carbonitriles Synthesis under Visible Light Irradiation. Catalysts. 2022; 12(12):1593. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12121593

Chicago/Turabian Style

Arunachalapandi, Murugan, Thangapandi Chellapandi, Gunabalan Madhumitha, Ravichandran Manjupriya, Kumar Aravindraj, and Selvaraj Mohana Roopan. 2022. "Direct Z-Scheme g-C3N5/Cu3TiO4 Heterojunction Enhanced Photocatalytic Performance of Chromene-3-Carbonitriles Synthesis under Visible Light Irradiation" Catalysts 12, no. 12: 1593. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12121593

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