Next Article in Journal
A Comparative Study of the Antiviral Properties of Thermally Sprayed Coatings against Human Coronavirus HCoV-229E
Next Article in Special Issue
The Comparison of Metal Doped TiO2 Photocatalytic Active Fabrics under Sunlight for Waste Water Treatment Applications
Previous Article in Journal
Comparison of Photocatalytic Biocidal Activity of TiO2, ZnO and Au/ZnO on Escherichia coli and on Aspergillus niger under Light Intensity Close to Real-Life Conditions
Previous Article in Special Issue
Effect of Calcination Temperature of SiO2/TiO2 Photocatalysts on UV-VIS and VIS Removal Efficiency of Color Contaminants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Stable Photocatalytic Dry and Bi-Reforming of Methane with the Role of a Hole Scavenger for Syngas Production over a Defective Co-Doped g-C3N4 Nanotexture

1
Chemical and Petroleum Engineering Department, UAE University, Al Ain P.O. Box 15551, United Arab Emirates
2
Department of Chemical Engineering, Balochistan University of Information Technology, Engineering and Management Sciences (BUITEMS), Quetta P.O. Box 87100, Pakistan
3
Center for Refining & Advanced Chemicals, King Fahd University of Petroleum & Minerals, Dhahran P.O. Box 31261, Saudi Arabia
4
Department of Chemistry, Maharshi Dayanand University, Rohtak 124001, India
5
Department of Electrical Engineering, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
6
Department of Physics and Astronomy, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(7), 1140; https://doi.org/10.3390/catal13071140
Submission received: 22 June 2023 / Revised: 13 July 2023 / Accepted: 20 July 2023 / Published: 22 July 2023

Abstract

:
Photocatalytic reduction of CO2 with CH4 through the dry reforming of methane (DRM) is an attractive approach to recycling greenhouse gases into valuable chemicals and fuels; however, this process is quite challenging. Although there is growing interest in designing efficient photocatalysts, they are less stable, and have lower photoactivity when employed for DRM reactions. Herein, we developed a noble metal-free hierarchical graphitic carbon nitride (HC3N4) loaded with cobalt (Co) for highly efficient and stable photocatalytic dry reforming of methane to produce synthesis gases (CO and H2). The performance of the newly designed Co/HC3N4 composite was tested for different reforming systems such as the dry reforming of methane, bi-reforming of methane (BRM) and reforming of CO2 with methanol–water. The performance of HC3N4 was much higher compared to bulk g-C3N4, whereas Co/HC3N4 was found to be promising for higher charge carrier separation and visible light absorption. The yield of CO and H2 with HC3N4 was 1.85- and 1.81-fold higher than when using g-C3N4 due to higher charge carrier separation. The optimized 2% Co/HC3N4 produces CO and H2 at an evolution rate of 555 and 41.2 µmol g−1 h−1, which was 18.28- and 1.74-fold more than using HC3N4 during photocatalytic dry reforming of methane (DRM), with a CH4/CO2 feed ratio of 1.0. This significantly enhanced photocatalytic CO and H2 evolution during DRM was due to efficient charge carrier separation in the presence of Co. The CH4/CO2 feed ratio was further investigated, and a 2:1 ratio was best for CO production. In contrast, the highest H2 was produced with a 1:1 feed ratio due to the competitive adsorption of the reactants over the catalyst surface. The performance of the composite was further investigated for bi-reforming methane and methanol. Using photocatalytic CO2 reduction with CH4/H2O, the production of CO and H2 was reduced, whereas significantly higher CO and H2 evolved using the BRM process involving methanol. Using methanol with CO2 and H2O, 10.77- and 1.39-fold more H2 and CO efficiency was achieved than when using dry reforming of methane. The composite was also very stable for continuous synthesis gas production during DRM in consecutive cycles. Thus, a co-assisted g-C3N4 nanotexture is promising for promoting photocatalytic activity and can be further explored in other solar energy applications.

1. Introduction

The industrialization and rapid growth of the economy has resulted in an increase in greenhouse gases, especially carbon dioxide (CO2) and methane (CH4), in the atmosphere. Both CO2 and CH4 have become the main causes of climate change and global warming [1,2]. Among other methods, the direct conversion of CO2 into valuable chemicals via catalytic reforming processes is intriguing and promising [3]. Through the dry reforming of methane (DRM) or catalytic CO2 reforming of CH4, CO2 can be used with methane to recycle both the gases into valuable chemicals and fuels [4]. DRM has a variety of exceptional benefits, including the reduction of both greenhouse emissions and the direct creation of syngas with an equimolar blend of CO and H2 [5]. However, DRM is more susceptible to coking, operates at higher reaction temperatures, and decreases catalyst stability [6,7].
As an alternative, using phototechnology is a potential method for the photocatalytic conversion of CO2 and CH4 to fuels under typical operating conditions. From this perspective, Shi et al. [8] reported the production of C2H6, CH3COOH, CH3COCH3, and CO products during the photocatalytic dry reforming of methane with a Cu/CdS-loaded TiO2/SiO2 composite. Similar to this, DRM has been studied over a Pt-loaded TiO2 catalyst, and under UV light irradiation, CO and H2 were shown to be the primary products [9]. Au/Rh-modified TNTs have recently been used in the photocatalytic dry reforming of methane to produce syngas [10]. Similarly, Rh/TiO2-B nanobelts were tested for the photocatalytic dry reforming of methane with the formation of CO and H2 UV and visible light irradiation [11]. Recently, we studied Ti3C2-supported TiO2 for the photocatalytic dry reforming of methane for synthesis gas production under UV-visible light irradiation [12]. Obviously, TiO2 has been extensively studied, but its poor activity is linked to the quick recombination rate of photo-generated charges, and it is only active under UV light [13]. Therefore, efficient and low-cost photocatalysts that are functional under solar energy are in high demand.
Among noble metal-free semiconductors, graphitic carbon nitride (g-C3N4) is becoming more important to scientists due to its several benefits, including affordability, improved potential for negative reduction, chemical and thermal stability, and its ability to be utilized under solar energy [14,15]. Its smaller specific surface area, compact layered structure, and faster rate of photoinduced charge carrier recombination are some of its key disadvantages [16,17]. To maximize photocatalytic efficiency, a number of enhancing techniques, including loading with metals and semiconductor coupling, have been investigated [18,19,20]. Limited reports are available on using g-C3N4 for the photocatalytic dry reforming of methane. Previously, we reported on a GO/g-C3N4 composite for use in the photocatalytic DRM process; photocatalytic efficiency was increased due to efficient charge carrier separation [21]. As cocatalysts, many noble metals, such as Pt, Au, Ag, and Ru, are frequently used to form and transport charge carriers to enhance g-C3N4 photocatalytic efficiency [22,23]. However, due to their high price and restricted availability, noble metals’ application for commercial hydrogen production is severely constrained. Due to their low cost, earth abundance, and reducibility, transition metals, including Ni, Co, Mn, and Fe, are seen as viable replacements for noble metals.
Cobalt (Co) is currently seen to be particularly promising due to numerous important properties like loosely bounded d-electrons, active sites, and effective electron trapping [24,25]. In addition to that, cobalt oxide has a narrow band gap, high surface area, exhibits strong light absorption, and has excellent electrical conductivity [26,27]. More carbon atoms can be attached to the active sites of the photocatalyst, increasing its stability and preventing the production of coke [28]. For instance, we studied hydrogen generation using Ni and Co loaded over TiO2, and found promising results and stability during the steam reforming of phenol to produce hydrogen [29]. Co-loaded TiO2 was found to have increased H2 evolution because it reduces overpotential and makes photoinduced electron transfer easier [30]. Another study found that loading g-C3N4 with a CoP cocatalyst free of noble metals increased the efficiency of H2 evolution [31]. To encourage solar energy-assisted hydrogen synthesis, additional research into cobalt with noble metal-free semiconductors is in high demand.
In this study, a facile approach to developing highly efficient and stable Co-doped g-C3N4 composites for the photocatalytic dry reforming of methane has been investigated. A hierarchical structured g-C3N4 (HC3N4) was synthesized through the controlled decomposition of melamine and urea and was loaded with Co using an impregnation approach. The role of structure and morphology was extensively investigated using a variety of characterization approaches with a variety of experimental validations. The performance evaluation was conducted based on synthesis gas (CO, H2) production during the photocatalytic dry reforming of methane (DRM) and bi-reforming with CH4/H2O and CH3OH/H2O systems. The effects of different CH4/CO2 feed ratios were investigated to understand the role of reactants’ attachment to catalysts’ active sites. Comparatively, Co/HC3N4 was found to be more promising for CO production compared to H2, and trends were consistent in all types of photocatalyst. Their stability was further investigated in cyclic runs, and catalysts were further characterized after the reactions. A reaction mechanism is further proposed to understand the role of catalysts and reactants in photoactivity and products’ selectivity.

2. Results and Discussion

2.1. Characterization of Materials

The XRD results for the samples of g-C3N4, HC3N4, and Co/HC3N4 are shown in Figure 1a. For g-C3N4, two diffraction peaks at 2θ of 12.84° and 27.42°, indexed to the (100) and (002) planes, respectively, were produced. The peak at the 12.84° plane is accounted for by the formation of tri-s-triazine chains, whereas another peak at 27.42° is attributed to aromatic rings in the fundamental in-plane periodic structure, as reported previously [32]. There was no change in the crystalline structure and peak position for the HC3N4, which confirms the successful synthesis of the hierarchical structured g-C3N4. Furthermore, crystalline planes were not minimally altered when Co was loaded over HC3N4, confirming the successful synthesis of a Co-loaded HC3N4 composite.
With the help of UV-vis DRS, the optical characteristics of g-C3N4, HC3N4, and Co/HC3N4 were further investigated, and the results are displayed in Figure 1b. There was little difference in the UV-visible light absorption spectra between g-C3N4 and HC3N4, definitely due to the crystalline structure remaining unchanged. When Co was loaded onto HC3N4, light absorption was somewhat boosted towards the visible region. Band gap energy values for g-C3N4, HC3N4 and Co/HC3N4 were estimated as 2.87, 2.89 and 2.86, respectively. This demonstrates that the influence of co-loading on the change in the band gap energy of HC3N4 was negligible. Recently, a g-C3N4 compound based on Co3O4 was created, and cobalt was found to increase visible light absorption [26].
Through PL analysis, the effectiveness of g-C3N4, HC3N4, and Co/HC3N4 was further studied, and the results are shown in Figure 1c. With pure g-C3N4, more charge carrier recombination led to a higher PL intensity. However, HC3N4 showed a lower peak intensity than g-C3N4, which is explained by less charge recombination. This demonstrates that hierarchical structure has a greater potential to lengthen the life of holes and electrons. When Co was loaded onto HC3N4, the lowest PL intensity was achieved due to efficient charge separation in the presence of cobalt. This shows that cobalt not only helps to increase active sites but also functions as a sink to capture and transport carriers of photoinduced charge. In many reports, enhanced charge separation in the presence of metals has been reported. For example, g-C3N4 PL intensity was significantly reduced with Co-loading [26].
FESEM was used to examine the structure and morphology of g-C3N4, HC3N4, and Co/HC3N4, and the results are shown in Figure 2. When melamine was thermally treated in an environment of airtight packed non-uniform sheets of g-C3N4, an uneven structure was obtained, as shown in Figure 2a. Figure 2b shows the morphology of hierarchical and exfoliated sheets of g-C3N4 (HC3N4) produced during the thermal decomposition of melamine with urea. Obviously, gaps are presented between the sheets, which would be useful for the penetration of more light irradiation, in addition to promoting the sorption process. When Co was loaded onto HC3N4, there was not much difference in the morphology of the pure HC3N4 and Co/HC3N4 samples. This reveals that Co has no effect in altering morphology; however, it was deposited over the layered structure during the impregnation method. The EDX mapping analysis further confirms the uniform distribution of Co over the HC3N4 surface, as shown in Figure 2d. The color images in Figure 2e–h further confirm the presence of Co, C, N, and O over the entire surface. All these findings confirm the successful fabrication of Co-loaded HC3N4 using a facile method.
The elemental state and composition of Co/HC3N4 were examined using X-ray photoelectron spectroscopy (XPS), and the results are presented in Figure 3. Figure 3a displays the N1s spectra with binding energies of 397.5, 399.0 and 401.5 eV, which are associated with C–N=C, N–(C)3, and C-NH2, respectively [33]. The high-resolution C 1s spectra in Figure 3b has binding energies of 284.6 and 288.1 eV that are attributed to C–C and N–C=N bonds in the sp2-hybridized aromatic ring units, respectively [34]. Four peaks in the Co 2p resolution spectrum in Figure 3c were located at binding energies of 780.8, 787.2, 797.3, and 802.3 eV. The binding energies assigned to the spin orbits Co 2p 3/2 and Co 2p1/2, respectively, were 780.6 and 797.3 eV and indexed to Co2+. The presence of cobalt in the oxide state is further supported by the two satellite peaks at 787.2 and 802.3 eV. All these results point to the successful loading of Co onto HC3N4 and to the presence of trapping and transporting charge carriers.

2.2. Photocatalytic Dry Reforming of Methane

To confirm that the results obtained during the photocatalytic reaction were solely obtained during the photocatalytic process, the photocatalytic activity of the synthesized samples was first assessed in a control experiment. To make sure that no product gases (CO, H2) were produced from the organic residues, the blank experiments were carried out utilizing N2 gas in the photoreactor under light illumination without any feed gases (CO2, CH4). The measurement of product gases (CO, H2) using an experiment without a photocatalyst was also carried out. This demonstrates that the photocatalytic reaction products were generated via a photocatalytic DRM process when light irradiation was present.

2.2.1. The Effect of Co-Loading on HC3N4 Photoactivity

The performances of g-C3N4 and HC3N4 were initially investigated for photocatalytic CO2 reduction with CH4 through a DRM reaction for the CO and H2 production, and their results are presented in Figure 4. CO and H2 yield rates of 18 and 14.78 µmol g−1 were produced with g-C3N4 and were increased to 33.35 and 26.8 µmol g−1, respectively, in the presence of HC3N4. This reveals that the hierarchical structure of g-C3N4 has an obvious effect on promoting photocatalytic activity. This augmented photoactivity was due to more light penetration in the nanosheets and its efficient interaction with reactants enabling more production and utilization of photoinduced charge carriers.
Figure 4 also shows the results of a photocatalytic dry reformation of methane (DRM) reaction that uses g-C3N4, HC3N4, and Co/HC3N4 samples as photocatalysts to produce syngas (CO and H2). In all types of photocatalyst, CO was the main product during the DRM process, and its yield rate was significantly increased with Co-loading onto HC3N4. Using 1% Co, a CO yield of 717.2 µmol g−1 was produced, which is 21.5- and 39.85-fold more than was produced with pure HC3N4 and g-C3N4 samples. The highest CO yield rate of 1172.86 µmol g−1 was obtained, which was 1.42- to 65-fold higher than using 3 and 1% Co-loading and pure g-C3N4 and HC3N4 catalysts. The trends for H2 production were similar to the CO, wherein the yield rate was much lower. Using 2% Co/HC3N4, the highest H2 yield of 92.36 µmol g−1 was produced, which is 1.49-, 2.1-, 3.5-, and 6.25-fold higher than 3% Co/HC3N4, 1% Co/C3N4, HC3N4 and g-C3N4 samples, respectively. Due to the efficient separation of charge carriers in the presence of Co, the amount of CO and H2 produced during the DRM reaction significantly increased [26]. However, any further increase in Co-loading beyond 2% Co decreases photocatalytic activity of HC3N4 due to the creation of charge recombination centers, which lowers the charge separation efficiency. Previously, photocatalytic H2 production with Co3O4/ZnIn2S4 was conducted, and it was observed that 3 wt. % Co3O4 produced the highest photocatalytic efficiency to maximize H2 production [35]. In a recent paper, we described the photocatalytic dry reforming of methane using a V2AlC MAX/g-C3N4 composite. The primary products obtained were CO and H2, with yield rates of 118.74 and 89.52 mol g−1 h−1, respectively [36]. The higher efficiency of V2AlC/g-C3N4 was due to efficient charge carrier separation during the photocatalytic process. In several other reports, the dry reforming of methane over various photocatalysts has been investigated, and the enhanced productivity was most probably due to the effective process of charge separation in the presence of cocatalysts and metals. Recently, with the use of Ti3C2 with TiO2, CO and H2, yield rates of 85 and 18 µmol g−1 h−1 were reported [12]. Similarly, CO and H2 yield rates of 99.8 and 5.6 µmol g−1 h−1 were observed during a DRM reaction with a GO/g-C3N4 composite. The significantly enhanced photocatalytic activity compared to pure materials was because of the effectual separation and utilization of photoinduced electrons and holes during the oxidation and reduction reactions.

2.2.2. The Effect of CH4/CO2 Feed Ratio for CO and H2 Production

Different CH4/CO2 feed ratios were chosen while keeping all other variables constant in order to examine the impact of feed ratios on the photocatalytic activity of 2 wt. % Co/HC3N4. Different feed ratios (CH4/CO2) and their relationships to CO and H2 evolution are depicted in Figure 5a. The chosen feed ratios for assessing the effect on CO and H2 evolution under visible light irradiation were CO2/CH4 1:1, 1:2, and 2:1. Using a CH4/CO2 input ratio of 1:1 affords the maximum rate of CO generation; however, a larger feed ratio reduces that production rate. The production of CO was reduced by adjusting to a feed ratio of 2:1. This revealed that a higher concentration of CH4 is more beneficial to enhancing the photocatalytic DRM reaction to maximize CO evolution. The highest CO production of 2703 µmol g−1 was achieved using a CH4/CO2 ratio of 2:1, which is 1.22- and 1.61-fold higher than using 1:1 and 2:1 feed ratios of CH4/CO2 under identical operating conditions. When increasing the CH4 concentration, there is more attachment of CH4 over the catalyst surface, resulting in greater production of protons that were effectively consumed in the reduction of CO2 to CO. In addition to this, there is the possibility of activating both a reversed water–gas shift (RWGS) reaction and reforming reactions, resulting in more production of CO [22,37]. Previously, it has been reported that CO2 attachment to the g-C3N4 surface is efficient to due to the basic nature of g-C3N4 compared to the acidic nature of CO2. Thus, by increasing the initial concentration of CH4, both the reactants may be effectively adsorbed for the DRM reaction to proceed. In photocatalysis, the attachment of both reactants is important for the oxidation and reduction reactions to proceed [19]. On the other hand, decreased CO evolution at much higher and lower concentrations may be due to the greater attachment of CO2 to the catalyst surface, the lower production of protons in the presence of a lesser amount of CH4, and inappropriate adsorption of both the reactants [38].
Figure 5b shows the effectiveness of Co/HC3N4 composites for the photocatalytic dry reforming of methane with various CH4/CO2 feed ratios to produce H2. Interestingly, varied CH4/CO2 feed ratios produce radically different outcomes for H2 generation compared with CO production. A large amount of H2 was produced using a CH4/CO2 feed ratio of 1.0 with an H2 yield of 164.80 µmol g−1, which was decreased to 118.67 and 95.27 µmol g−1 with the increase in the CH4/CO2 feed ratio. These results can be discussed based on different hypotheses, such as the reactants’ adsorption compositions and competitive reactions occurring over the photocatalyst surface. At a CH4/CO2 feed ratio of 1.0, there should be an effective attachment of both the reactants, resulting in a higher rate of oxidation and reduction reactions. Due to the lesser availability of protons, there wer fewer changes in the consumption of hydrogen to be used for the RWGS reaction; thus, a higher amount of H2 was produced. Likewise, when the feed ratio was increased further by increasing the CH4 concentration, it became possible for the coupling of methane to produce products other than hydrogen, resulting in a lower production of hydrogen [39,40]. Therefore, an optimized CH4/CO2 feed ratio is required to maximize the production of both CO and H2 during the photocatalytic dry reforming of methane reactions.
In several other works, the effects of CH4/CO2 feed ratios on product yield and selectivity with various types of photocatalyst have been reported. For instance, the effectiveness of GO/g-C3N4 was studied using various CH4/CO2 feed ratios. It was observed that the highest CO production was obtained with a CO2:CH4 ratio of 1.0, whereas H2 production was favored in a methane-rich feed mixture with a CH4:CO2 feed ratio of 2:1 [21]. Similarly, the effects of different feed ratios in DRM and BRM reactions on the performance of Ti3C2/g-C3N4/LDH for CO and H2 evolution have been investigated. Using a DRM process with a CO2:CH4 feed ratio of 2:1, the highest yield of CO was obtained; however, a CO2:CH4 feed ratio of 1:2 was conducive to enhancing H2 evolution. In the case of the BRM process, again, the highest CO was observed when a CO2:CH4 feed ratio of 2:1 with H2O was tested. Similar trends for H2 evolutions were observed, in which the maximum yield was obtained with a CO2:CH4 feed ratio of 1:2 with H2O [41].

2.2.3. Photocatalytic CO2 Bi-Reforming of Methane and Methanol

In photocatalytic CO2 reforming systems, the product yield and selectivity are entirely dependent on the type of feed mixture and its composition. Therefore, the performance and selectivity of Co/g-C3N4 was further investigated using various reforming systems such as the bi-reforming of methane (CO2 with CH4/H2O) and the bi-reforming of methanol (CO2 with CH3OH/H2O) under identical reaction conditions. To explore photocatalytic CO2 reduction through bi-reforming, equal concentrations of CO2 and CH4 feed mixture (CH4/CO2 = 1.0) were passed through a water saturator to carry moisture before entering the reactor. To explore the bi-reforming of CO2 with methanol as a hole scavenger, 10 vol. % methanol/water was first prepared and used to saturate CO2 gas with methanol/water vapors.
Photocatalytic CO2 reduction through the bi-reforming of methane (CH4/H2O) for the production of CO and H2 is shown in Figure 6a. Over the course of the entire irradiation period, it was evidently possible to produce continuous CO and H2; nevertheless, CO production was more significant than H2 production, and these results are consistent with those of the DRM reaction. Maximum CO and H2 evolutions of 689 and 95.8 µmol g−1 were attained over 2% Co/HC3N4 after 4 h of irradiation time. By comparing the results with the DRM process, it was detected that both the CO and H2 production were decreased by introducing water with CH4/CO2 during the BRM process of CO2 reduction. These results could be explained based on several hypotheses due to the complex photocatalysis process, which involves competitive oxidation and reduction reactions. When using feed mixtures of CO2 and CH4, they have equal potential to attach to the photocatalyst surface, promoting both oxidation and reduction reactions. CO2 is consumed to produce CO through the use of electrons and protons, whereas CH4 is oxidized to produce protons. On the other hand, when the CO2/CH4 mixture was used with water, there may be more attachment of water molecules over the Co/g-C3N4 surface and fewer active sites available for the attachment of CO2 molecules to be consumed in the reduction reaction. Therefore, although the oxidation reaction will be efficient, the reduction reaction will be lacking, resulting in a lower overall photocatalytic efficiency. Thus, efficient attachment of both the reactants is necessary to maximize oxidation and reduction reactions for the production of CO and H2 over a Co/HC3N4 photocatalyst.
The photocatalytic reduction of CO2 with methanol and water solution through BRM was further investigated, and the results are presented in Figure 6b. The main products obtained were CO and H2, with smaller amount of CH4. Using both H2O and CH3OH as sacrificial reagents, continuous generation of CO, CH4, and H2 was achieved throughout the irradiation period. Similar to DRM and BRM process, CO2 reduction with methanol and water was conducive to producing the highest yield of CO and H2 over the Co/HC3N4 composite photocatalyst. This BRM system with methanol and water gave the highest yields of 3086, 1776 and 20.96 µmol g−1 for CO, H2 and CH4, respectively, after four hours of irradiation time. The amount of CO produced with the methanol–water system was 1.39 and 4.42 times higher than using the CO2-CH4 and CO2-CH4-H2O reforming systems. Similarly, the yield of H2 was 10.8- and 18.5-fold more than when it evolved through the involvement of the DRM and BRM reactions. All these findings reveal that methanol as a sacrificial reagent is more able to produce significant amounts of H2 and CO, while keeping all other parameters constant. This further demonstrates that HC3N4 in any type of reforming system is likely to boost CO generation as the primary product, and produce an appreciable amount of H2 due to the suitable reduction potential of CO2/CO (−0.50 eV) and H+/H2 (−0.41 eV) with the CB of g-C3N4 (−1.30 eV).
This demonstrates that for the conversion of CO2 to CO and H2, methanol works better than Co/HC3N4 as a sacrificial reagent. The increased attachment of methanol to the surface of g-C3N4 was the cause of its photocatalytic activity. Sacrificial reagents, in general, trap holes and are helpful in producing electrons, resulting in an abundance of protons and electrons to produce H2. As a result, the photoactivity of CO and H2 production was higher with OH-based sacrificial reagents such as CH3OH. This is because methanol and water have more potential to adhere to the catalytic surface than CH4. Due to the unique surface functional group of g-C3N4, methanol will more effectively adsorb over the Co/HC3N4 surface. This facilitates an effective CO2 photoreduction process by facilitating CH3OH’s attachment to the g-C3N4 surface [15].
In various reports, the effect of different sacrificial reagents has been investigated. For example, the effect of different reducing agents during CO2 reforming reactions (carried out to examine the performance of Ti3C2/TiO2) was reported. It was observed that using a CO2-H2O feed mixture, the highest amount of CO was produced, whereas the bi-reforming of methane (CO2-CH4-H2O) was more promising for increasing the H2 evolution rate. These results were possibly due to the acidic characteristic of the catalysts. The favorable H2 production achieved with BRM reforming systems was possibly due to higher CH4 and H2O adsorption over the catalyst’s active sites, which causes hydrogen-rich synthesis gas production [12]. Previously, the performances of GO/g-C3N4 composites under different reforming systems (such as the dry reforming of methane and the bi-reforming of methane) have been investigated. It was found that DRM is the most efficient method of producing more CO and H2, compared to BRM and CO2 reduction with H2O [21].

2.2.4. Performance Analysis

A performance comparison of the Co/HC3N4 composite under different operating conditions and reforming systems is summarized in Table 1. The performance of the composite was investigated using the dry reforming of methane (DRM), the bi-reforming of methane (BRM) and CO2 reforming with a methanol–water mixture in a fixed-bed photoreactor system. Using all types of reforming systems, CO was identified as the main product. However, the production of H2 was dependent on the kind of reforming systems, whereas the highest H2 yield was produced using methanol as the sacrificial reagent.
Using HC3N4, CO and H2 production rates of 30.36 and 23 µmol g−1 h−1 were produced, due to modifying structure to exfoliated nanosheets. With loading optimized 2% Co with HC3N4, CO and H2 yield rates of 555 and 41.2 µmol g−1 h−1 were produced during DRM process with a CH4/CO2 feed ratio of 1.0. The yield of CO and H2 over Co/HC3N4 was 18.28- and 1.74-fold higher than when using HC3N4, respectively. The production of CO was increased to 676 µmol g−1 h−1 with a CH4/CO2 feed ratio of 2:1, whereas H2 production was decreased. The CO production was further increased to 771 µmol g−1 h−1 when methanol was used as the sacrificial reagent. Similar to this, the highest H2 of 444 µmol g−1 h−1 was produced over 2% Co/HC3N4 with methanol as the sacrificial reagent, which is 10.78, 14.95- and 18.67-fold higher than that produced when using CH4/CO2 feed ratios of 1:1, 2:1, and 2:1, respectively. More importantly, in the bi-reforming of methane (CO2-CH2-H2O), the lowest photocatalytic efficiency was observed. All these findings point to the higher efficiency of the Co/HC3N4 composite in the conversion of stable CO2 and CH4 compounds to CO and H2 through a photocatalytic process.
Few reports are available on the photocatalytic dry reforming of methane under UV and visible light irradiations. In a recent work, a covalent organic framework (COF)-based composite was tested for the photocatalytic dry reforming of methane. CO and H2 yield rates of 56.18 and 1.54 µmol g−1 h−1, respectively, were obtained, which were significantly higher than those obtained with pure materials [42]. Similarly, we reported a V2AlC MAX/g-C3N4 composite for the photocatalytic dry reforming of methane, and the main products obtained were CO and H2, with yield rates of 118.74 and 89.52 µmol g−1 h−1, respectively [36]. In several other reports, the dry reforming of methane over various photocatalysts has been investigated. Recently, with the use of Ti3C2 with TiO2, CO and H2, yield rates of 85 and 18 µmol g−1 h−1 were reported [12]. Similarly, CO and H2 yield rates of 99.8 and 5.6 µmol g−1 h−1 were observed during a DRM reaction over a GO/g-C3N4 composite. By comparing the results with previous works, it can be seen that noble metal-free HC3N4 loaded with Co is a very promising material for maximizing the dry reforming of methane to produce synthesis gas.

2.3. Stability Analysis

Given that the catalyst deactivates during the dry reformation of methane, it is imperative to conduct stability research into the synthesized 2% Co/HC3N4 composite for syngas (CO and H2) through a DRM reaction. Figure 7a shows the findings of a stability test that was carried out in this work to assess how well the synthesized composite performed in producing syngas (CO and H2) over several cycles, when exposed to visible light. The catalyst was uniformly dispersed throughout the reactor for the stability test, and the CH4/CO2, input feed ratio, was maintained at 1.0 for each cycle. Evidently, consistent CO and H2 generation was achieved throughout each of the stability tests. For CO yield, it was increased in the second cyclic run, and this was made possible by the reductive characteristics of cobalt under light irradiation. However, in the third cyclic run, still, higher and continuous production of CO was obtained compared to the first cyclic run. Recently, a covalent organic framework (COF)-based composite was tested and reported to increase CO evolution in the second cyclic run during the photocatalytic dry reforming of methane [42]. On the other hand, during the whole irradiation and throughout all of the cyclic runs, H2 was continuously and steadily produced. This demonstrated that the catalyst remained stable throughout the DRM process for the evolution of CO and H2. After three cycles, it is evident from the CO and H2 evolution data that the composite performs extremely well as a photocatalyst during the DRM process when exposed to visible light. Previously, we investigated DRM over V2AlC/g-C3N4; however, therein, the photocatalytic activity for CO and H2 declined over the irradiation time [36].
XRD analysis was used to further compare the 2% Co/HC3N4 composites before and after the reaction, and the results are displayed in Figure 7b. The fresh and used samples both had similar peaks, proving that the peak position had not altered. The composite obtained after the reaction showed no new peaks, demonstrating the material’s excellent stability and the absence of structural change or carbon generation. This demonstrates that using cobalt in combination with g-C3N4 rendered higher stability for the DRM process. Figure 7c shows the results of an FTIR study that was carried out to further evaluate the used photocatalyst before and after three cycles of the photocatalytic DRM reaction. The FTIR spectra of the composite are nearly identical before and after the reaction. The FTIR spectra exhibit the same peaks before and after the reaction; no additional peaks are present, thereby confirming the higher stability of g-C3N4 when loaded with cobalt during the DRM process.

2.4. Proposed Mechanism

When it comes to improved photocatalytic activity, a photocatalyst’s performance is determined by effective charge separation and charge carrier mobility (e and h+). Similar to this, fast charge separation and consideration of the reduction potential of the products (CO and H2) in relation to the semiconductors’ conduction band locations are also required for photocatalytic reduction via DRM and BRM processes. Additionally, other factors including light absorption capacity, reactant activation, and their adsorption over the catalyst’s surface are essential factors required for an effective photocatalysis process. Furthermore, increased exposed surface area encourages more light absorption, which produces a lot of charge to boost the reduction and oxidation reactions.
Due to compact nanosheets with bulk structures, the g-C3N4 performs charge separation less effectively, resulting in a lower amount of CO and H2 evolution. On the other hand, when g-C3N4 was exfoliated to produce a hierarchical structure, there were obvious gaps to enable efficient light penetration with defects, itself then enabling efficient charge carrier separation. The layered structure of HC3N4 allows strong light penetration with efficient charge carrier separation, resulting in higher photocatalytic efficiency.
Figure 8 provides a schematic representation of the suggested mechanism for the dry reforming of methane over a Co/HC3N4 composite. As was previously mentioned, CO and H2 were the primary products of the photocatalytic DRM reaction over the Co/HC3N4 composite, which were produced through the involvement of two electrons. This suggests that numerous charge carriers and effective charge carrier separation are required for the increased and selective production of syngas (CO and H2). In order to facilitate quick charge separation, Co was found to be a promising vehicle to trap and transport photoinduced charge carriers. The reactions involved during the oxidation and reduction reactions are discussed in Equations (1)–(11) [12,43].
Oxidation reactions:
C H 4 + h +     C H 3 + H +
C H 3 + h +     C H 2 + H +
C H 2 + h +     C H + H +
C H + h +   C + H +
Reduction reactions:
C O 2 + e     C O 2
C O 2 + 2 H + + e C O + H 2 O
C O 2 + 2 H + + 2 e     C O + H 2 O
Series reactions for CO and H2 formation:
C H 3 + H 2 O + 6 h +   C O + 5 H +
C H 2 + H 2 O + 5 h +   C O + 4 H +
C H + H 2 O + 4 h +   C O + 3 H +
C + H 2 O + 3 h +   C O + 2 H +
2 H + + 2 e H 2
During the photocatalytic dry reforming of methane, methane is oxidized by holes to produce intermediate ∙CHx−1 and protons (H+) through the involvement of the reactions in Equations (1)–(4). The efficiency of these reactions depends on the attachment of CH4 to the catalyst surface, and also on the availability of holes to proceed with the oxidation reaction. Due to this series of reactions (as detailed in Equations (1) and (4)), final products of H+ and C   o are produced. The protons produced are further used in a series of reactions to produce CO during the reduction reaction. During the reduction reaction, CO2 is reduced with electrons to produce intermediate C O 2 through the reaction in Equation (5) with the involvement of electrons. The intermediate products ( C O 2 ) and CO2 can be converted to CO and water molecules through the involvement of two protons, as shown in Equations (6) and (7). The CH4 intermediates, with the involvement of water molecules and holes, produce CO and protons, as discussed in Equations (8)–(11). However, in all these reactions, the important factor is the utilization of Co with water molecules and holes to produce CO and protons (Equation (11)). Finally, H2 is produced through the use of two electrons, as shown in Equation (12) [22,37]. The production of CO occurred through several parallel reactions with the involvement of holes, protons, and electrons over the highly efficient Co/HC3N4 composite photocatalyst. Thus, the production of CO was higher compared to hydrogen production during the dry reforming of methane over the Co/HC3N4 composite photocatalyst.

3. Experimental

3.1. Synthesis of Hierarchical Graphitic Carbon Nitride (HC3N4)

Hierarchical and bulk samples of carbon nitrides were created using melamine as the precursor, using the hydrothermal method. g-C3N4 was synthesized using only melamine and heating for two hours at 550 °C, as reported previously [44]. For the synthesis of HC3N4, equal amounts of melamine and urea were heated at 550 °C with total time of 2 h in an air atmosphere. The final product obtained was given the name (HC3N4).

3.2. Synthesis of Co-Doped g-C3N4 (Co/HC3N4)

The co-loaded HC3N4 nanotextures were produced using an impregnation method. A predetermined amount of HC3N4 was dispersed in methanol, and then varied amounts of cobalt nitrates dissolved in methanol were added to the above suspension under continuous stirring. The slurry was agitated for two hours before being dried overnight at 100 °C. The finished product was denoted Co/HC3N4. Different Co-loading amounts (1, 2 and 3 wt. %) were prepared using various cobalt nitrate loadings with HC3N4; these were then labeled 1% Co/HC3N4, 2% Co/HC3N4, and 3% Co/HC3N4, respectively.

3.3. Characterization

The material’s structure, crystallinity, morphology, element dispersion, light absorption, and ability to separate charges were all investigated using a range of instruments. An XRD analysis was performed to determine the crystalline structure using the Bruker Advance D8 diffractometer. Morphology analysis was carried out using a Hitachi SU8020 SEM (Scanning electron microscope). We used a PerkinElmer equipment to obtain FTIR spectra. XPS analysis was carried out using an Axis Ultra DLD instrument. Using the UV-3600 Plus Spectrometer, the band gap energy and light absorption were estimated. A PL analysis was then carried out utilizing a spectrometer from HORIBA Scientific with a laser of wavelength 325 nm. A Raman analysis was performed using HORIBA Scientific with a laser of wavelength 532 nm.

3.4. Photocatalytic Activity Test

A photoactivity test to assess the performance of synthesized photocatalysts was conducted in a stainless steel photoreactor. The light source used was a 200 W Hg lamp with light intensity 100 mW cm−2, which was equipped with a cooling fan to remove lamp heat. Typically, a 150 mg powder photocatalyst was uniformly dispersed within the stainless-steel chamber, equipped with a quartz glass. Mass flow controls were used to control the flow rate of CO2 and CH4. In order to saturate the catalyst and purge the reactor before the experiment began, the feed mixture (CO2 and CH4) was continually pumped through the reactor for 30 min at a total flow rate of 20 mL/min and a CO2/CH4 molar feed ratio of 1.0. The effects of different CH4/CO2 feed ratios were further investigated to understand the role of feed composition on the photocatalytic performance. For this purpose, different feed mixtures of CH4/CO2 were used, while keeping a total flow rate of 20 mL/min. During the photocatalytic bi-reforming of methane, the mixture of CO2/CH4 was passed through the water saturator before entering the reactor. To further investigate the effect of the reducing agent through the bi-reforming of CO2 with methanol/water, CO2 was passed through a 10% methanol/water solution at a total flow rate of 20 mL/min. A gas chromatographer (GC-Agilent Technologies 6890 N, USA) connected to a thermal conductivity detector (TCD) and a flame-ionized detector (FID) was used to evaluate the products.

4. Conclusions

In conclusion, self-assembly was used to produce Co/HC3N4, which offers a potential method of carrying out the photocatalytic dry reforming of methane by utilizing both CH4 and CO2. The DRM was tested using a variety of catalyst materials, including pure and composite samples, to examine their photocatalytic activity. In all types of photocatalysts, CO was identified as the main product, with a lower amount of H2 during CO2 reduction with CH4 through DRM and with CH4/H2O through the BRM process. Excellent photocatalytic activity for the formation of syngas was achieved by the Co/HC3N4, which was much higher than that achieved when using pure HC3N4 and g-C3N4 samples. When the CH4/CO2 feed ratio was altered, the yields of CO and H2 also varied. A CH4/CO2 feed ratio of 2:1 was favorable for maximizing CO evolution, whereas a CH4/CO2 feed ratio of 1 was more suitable for obtaining the highest H2 productivity. Among the different reforming systems, the bi-reforming of CO2 with methanol and water was the most efficient system for maximizing the production of both CO and H2 due to the production of more electrons and protons. The cobalt-assisted HC3N4 was also promising in its demonstration of higher photostability in multiple cycles for the continuous production of CO and H2 during the DRM process. In general, excellent interfacial contact between Co and HC3N4 enables promising charge carrier separation with higher photostability, and is promising for other energy-based and environmental applications.

Author Contributions

Writing—original draft preparation, review and editing, M.T.; data curation, A.A.K.; data analysis, A.B.; reviewing and editing, N.K.; software, M.S.; resources, A.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by United Arab Emirates University (UAEU) research under fund number 12N097.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khan, A.A.; Tahir, M. Recent advancements in engineering approach towards design of photo-reactors for selective photocatalytic CO2 reduction to renewable fuels. J. CO2 Util. 2019, 29, 205–239. [Google Scholar] [CrossRef]
  2. Waris, M.; Ra, H.; Yoon, S.; Kim, M.-J.; Lee, K. Efficient and Stable Ni/SBA-15 Catalyst for Dry Reforming of Methane: Effect of Citric Acid Concentration. Catalysts 2023, 13, 916. [Google Scholar] [CrossRef]
  3. Khoja, A.H.; Tahir, M.; Amin, N.A.S. Dry reforming of methane using different dielectric materials and DBD plasma reactor configurations. Energy Convers. Manag. 2017, 144, 262–274. [Google Scholar] [CrossRef]
  4. Moral, A.; Reyero, I.; Alfaro, C.; Bimbela, F.; Gandia, L.M. Syngas production by means of biogas catalytic partial oxidation and dry reforming using Rh-based catalysts. Catal. Today 2018, 299, 280–288. [Google Scholar] [CrossRef]
  5. Goscianska, J.; Pietrzak, R.; Matos, J. Catalytic performance of ordered mesoporous carbons modified with lanthanides in dry methane reforming. Catal. Today 2018, 301, 204–216. [Google Scholar] [CrossRef]
  6. Huang, N.; Su, T.; Qin, Z.; Ji, H. Nickel Supported on Multilayer Vanadium Carbide for Dry Reforming of Methane. ChemistrySelect 2022, 7, e202203873. [Google Scholar] [CrossRef]
  7. Wu, S.; Li, Y.; Hu, Q.; Wu, J.; Zhang, Q. Photothermocatalytic Dry Reforming of Methane for Efficient CO2 Reduction and Solar Energy Storage. ACS Sustain. Chem. Eng. 2021, 9, 11635–11651. [Google Scholar] [CrossRef]
  8. Shi, D.; Feng, Y.; Zhong, S. Photocatalytic conversion of CH4 and CO2 to oxygenated compounds over Cu/CdS-TiO2/SiO2 catalyst. Catal. Today 2004, 98, 505–509. [Google Scholar] [CrossRef]
  9. Han, B.; Wei, W.; Chang, L.; Cheng, P.; Hu, Y.H. Efficient Visible Light Photocatalytic CO2 Reforming of CH4. ACS Catal. 2015, 6, 494–497. [Google Scholar] [CrossRef]
  10. László, B.; Baán, K.; Varga, E.; Oszkó, A.; Erdőhelyi, A.; Kónya, Z.; Kiss, J. Photo-induced reactions in the CO2-methane system on titanate nanotubes modified with Au and Rh nanoparticles. Appl. Catal. B Environ. 2016, 199, 473–484. [Google Scholar] [CrossRef]
  11. Kushida, M.; Yamaguchi, A.; Miyauchi, M. Photocatalytic dry reforming of methane by rhodium supported monoclinic TiO2-B nanobelts. J. Energy Chem. 2022, 71, 562–571. [Google Scholar] [CrossRef]
  12. Khan, A.A.; Tahir, M.; Zakaria, Z.Y. Synergistic effect of anatase/rutile TiO2 with exfoliated Ti3C2TR MXene multilayers composite for enhanced CO2 photoreduction via dry and bi-reforming of methane under UV–visible light. J. Environ. Chem. Eng. 2021, 9, 105244. [Google Scholar] [CrossRef]
  13. Shehzad, N.; Tahir, M.; Johari, K.; Murugesan, T.; Hussain, M. A critical review on TiO2 based photocatalytic CO2 reduction system: Strategies to improve efficiency. J. CO2 Util. 2018, 26, 98–122. [Google Scholar] [CrossRef]
  14. He, Y.; Jiang, G.; Jiang, Y.; Tang, L.; Zhang, C.; Guo, W.; Li, F.; You, L.; Li, S.; Liu, X. In-situ nanoarchitectonics of noble-metal-free g-C3N4@C-Ni/Ni2P cocatalyst with core-shell structure for efficient photocatalytic H2 evolution. J. Alloys Compd. 2022, 895, 162583. [Google Scholar] [CrossRef]
  15. Tahir, M.; Tahir, B. In-situ growth of TiO2 imbedded Ti3C2TA nanosheets to construct PCN/Ti3C2TA MXenes 2D/3D heterojunction for efficient solar driven photocatalytic CO2 reduction towards CO and CH4 production. J. Colloid Interface Sci. 2021, 591, 20–37. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, Z.; Guo, F.; Sun, H.; Shi, Y.; Shi, W. Well-designed three-dimensional hierarchical hollow tubular g-C3N4/ZnIn2S4 nanosheets heterostructure for achieving efficient visible-light photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2022, 607 Pt 2, 1391–1401. [Google Scholar] [CrossRef]
  17. Tahir, M. Investigating the Influential Effect of Etchant Time in Constructing 2D/2D HCN/MXene Heterojunction with Controlled Growth of TiO2 NPs for Stimulating Photocatalytic H2 Production. Energy Fuels 2021, 35, 6807–6822. [Google Scholar] [CrossRef]
  18. Qi, H.; Wang, C.; Shen, L.; Wang, H.; Lian, Y.; Zhang, H.; Ma, H.; Zhang, Y.; Zhang, J.Z. α-NiS/g-C3N4 Nanocomposites for Photocatalytic Hydrogen Evolution and Degradation of Tetracycline Hydrochloride. Catalysts 2023, 13, 983. [Google Scholar] [CrossRef]
  19. Muhammad, A.; Tahir, M.; Al-Shahrani, S.S.; Mahmood Ali, A.; Rather, S.U. Template free synthesis of graphitic carbon nitride nanotubes mediated by lanthanum (La/g-CNT) for selective photocatalytic CO2 reduction via dry reforming of methane (DRM) to fuels. Appl. Surf. Sci. 2020, 504, 144177. [Google Scholar] [CrossRef]
  20. Tahir, M.; Tahir, B.; Zakaria, Z.Y.; Muhammad, A. Enhanced photocatalytic carbon dioxide reforming of methane to fuels over nickel and montmorillonite supported TiO2 nanocomposite under UV-light using monolith photoreactor. J. Clean. Prod. 2019, 213, 451–461. [Google Scholar] [CrossRef]
  21. Ikreedeegh, R.R.; Tahir, M. Facile fabrication of well-designed 2D/2D porous g-C3N4–GO nanocomposite for photocatalytic methane reforming (DRM) with CO2 towards enhanced syngas production under visible light. Fuel 2021, 305, 121558. [Google Scholar] [CrossRef]
  22. Tahir, B.; Tahir, M.; Amin, N.A.S. Ag-La loaded protonated carbon nitrides nanotubes (pCNNT) with improved charge separation in a monolithic honeycomb photoreactor for enhanced bireforming of methane (BRM) to fuels. Appl. Catal. B 2019, 248, 167–183. [Google Scholar] [CrossRef]
  23. Tahir, B.; Tahir, M.; Amin, N.A.S. Photo-induced CO2 reduction by CH4/H2O to fuels over Cu-modified g-C3N4 nanorods under simulated solar energy. Appl. Surf. Sci. 2017, 419, 875–885. [Google Scholar] [CrossRef]
  24. Yang, P.; Wang, R.; Tao, H.; Zhang, Y.; Titirici, M.-M.; Wang, X. Cobalt Nitride Anchored on Nitrogen-Rich Carbons for Efficient Carbon Dioxide Reduction with Visible Light. Appl. Catal. B 2021, 280, 119454. [Google Scholar] [CrossRef]
  25. Fan, W.K.; Tahir, M. Recent advances on cobalt metal organic frameworks (MOFs) for photocatalytic CO2 reduction to renewable energy and fuels: A review on current progress and future directions. Energy Convers. Manag. 2022, 253, 115180. [Google Scholar] [CrossRef]
  26. Yang, X.; Chen, W.; Liu, Y.; Liang, B.; Hu, J.; Huang, H. Dual-Z-scheme Co3O4/CoO-CNNS heterojunction for facilitating charge transport toward improving photocatalytic hydrogen generation. J. Alloys Compd. 2023, 939, 168807. [Google Scholar] [CrossRef]
  27. Zhang, W.; Descorme, C.; Valverde, J.L.; Giroir-Fendler, A. Effect of Calcination Conditions on Co3O4 Catalysts in the Total Oxidation of Toluene and Propane. Catalysts 2023, 13, 992. [Google Scholar] [CrossRef]
  28. An, Z.; Wang, W.; Dong, S.; He, J. Well-distributed cobalt-based catalysts derived from layered double hydroxides for efficient selective hydrogenation of 5-hydroxymethyfurfural to 2,5-methylfuran. Catal. Today 2019, 319, 128–138. [Google Scholar] [CrossRef]
  29. Abbas, T.; Tahir, M. Tri-metallic Ni–Co modified reducible TiO2 nanocomposite for boosting H2 production through steam reforming of phenol. Int. J. Hydrogen Energy 2021, 46, 8932–8949. [Google Scholar] [CrossRef]
  30. Chen, W.; Liu, M.; Wang, Y.; Gao, L.; Dang, H.; Mao, L. Non-noble metal Co as active sites on TiO2 nanorod for promoting photocatalytic H2 production. Mater. Res. Bull. 2019, 116, 16–21. [Google Scholar] [CrossRef]
  31. Peng, S.; Huang, X.; Cao, Y.; Zuo, J.; Huang, Y.; Xu, J.; Li, Y. Modification of carbon nitride with noble-metal-free cobalt phosphide for effective photocatalytic hydrogen evolution. Appl. Surf. Sci. 2022, 584, 152610. [Google Scholar] [CrossRef]
  32. Ansari, H.M.; Wang, W.; Lei, L.; Bao, K.; Chang, X.; Raza, A.; Chen, Y.; Mehboob, A.; Zhong, Q.; Srivastava, A.; et al. Gas bubbling exfoliation strategy towards 3D g-C3N4 hierarchical architecture for superior photocatalytic H2 evolution. J. Alloys Compd. 2022, 919, 165794. [Google Scholar] [CrossRef]
  33. Sun, X.; Li, L.; Hu, T.; Lian, Q.; Liu, Y.; Hou, J.; Yu, Y.; Cao, Y.; Ma, F. In2S3/g-C3N4/CoZnAl-LDH composites with the lamellar dual S-scheme heterostructure and its enhanced photocatalytic performance. Colloids Surf. A Physicochem. Eng. Asp. 2023, 658, 130744. [Google Scholar] [CrossRef]
  34. Liang, S.; Sui, G.; Guo, D.; Luo, Z.; Xu, R.; Yao, H.; Li, J.; Wang, C. g-C3N4-wrapped nickel doped zinc oxide/carbon core-double shell microspheres for high-performance photocatalytic hydrogen production. J. Colloid Interface Sci. 2023, 635, 83–93. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Chen, D.; Li, N.; Xu, Q.; Li, H.; Lu, J. Fabricating 1D/2D Co3O4/ZnIn2S4 core–shell heterostructures with boosted charge transfer for photocatalytic hydrogen production. Appl. Surf. Sci. 2023, 610, 155272. [Google Scholar] [CrossRef]
  36. Madi, M.; Tahir, M. Fabricating V2AlC/g-C3N4 nanocomposite with MAX as electron moderator for promoting photocatalytic CO2-CH4 reforming to CO/H2. Int. J. Energy Res. 2022, 46, 7666–7685. [Google Scholar] [CrossRef]
  37. Tahir, B.; Tahir, M.; Amin, N.A.S. Tailoring performance of La-modified TiO2 nanocatalyst for continuous photocatalytic CO2 reforming of CH4 to fuels in the presence of H2O. Energy Convers. Manag. 2018, 159, 284–298. [Google Scholar] [CrossRef]
  38. Tahir, M.; Mansoor, R. Constructing a stable 2D Ti3AlC2 MnAXm cocatalyst-modified gC3N4/CoAl-LDH/Ti3AlC2 heterojunction for efficient dry and bireforming of methane for photocatalytic syngas production. J. Alloys Compd. 2023, 947, 169457. [Google Scholar] [CrossRef]
  39. Tahir, M.; Tahir, B.; Amin, N.A.S.; Muhammad, A. Photocatalytic CO2 methanation over NiO/In2O3 promoted TiO2 nanocatalysts using H2O and/or H2 reductants. Energy Convers. Manag. 2016, 119, 368–378. [Google Scholar] [CrossRef]
  40. Tahir, B.; Tahir, M.; Amin, N.A.S. Silver loaded protonated graphitic carbon nitride (Ag/pg-C3N4) nanosheets for stimulating CO2 reduction to fuels via photocatalytic bi-reforming of methane. Appl. Surf. Sci. 2019, 493, 18–31. [Google Scholar] [CrossRef]
  41. Ali Khan, A.; Tahir, M. Constructing S-Scheme Heterojunction of CoAlLa-LDH/g-C3N4 through Monolayer Ti3C2-MXene to Promote Photocatalytic CO2 Re-forming of Methane to Solar Fuels. ACS Appl. Energy Mater. 2021, 5, 784–806. [Google Scholar] [CrossRef]
  42. Yin, S.-Y.; Li, Z.; Hu, Y.; Luo, X.; Li, J. A novel metal-free porous covalent organic polymer for efficient room-temperature photocatalytic CO2 reduction via dry-reforming of methane. Green Energy Environ. 2023; in press. [Google Scholar]
  43. Tahir, M.; Tahir, B.; Amin, N.S. Photocatalytic CO2 reduction by CH4 over montmorillonite modified TiO2 nanocomposites in a continuous monolith photoreactor. Mater. Res. Bull. 2015, 63, 13–23. [Google Scholar] [CrossRef]
  44. Khan, A.A.; Tahir, M.; Mohamed, A.R. Constructing S-scheme heterojunction of carbon nitride nanorods (g-CNR) assisted trimetallic CoAlLa LDH nanosheets with electron and holes moderation for boosting photocatalytic CO2 reduction under solar energy. Chem. Eng. J. 2022, 433, 133693. [Google Scholar] [CrossRef]
Figure 1. (a) XRD analysis of g-C3N4, HC3N4, Co/HC3N4, (b) UV-vis DRS analysis of g-C3N4, HC3N4, Co/HC3N4, (c) PL analysis of g-C3N4, HC3N4, Co/HC3N4.
Figure 1. (a) XRD analysis of g-C3N4, HC3N4, Co/HC3N4, (b) UV-vis DRS analysis of g-C3N4, HC3N4, Co/HC3N4, (c) PL analysis of g-C3N4, HC3N4, Co/HC3N4.
Catalysts 13 01140 g001
Figure 2. SEM images of (a) g-C3N4, (b) HC3N4, (c) Co/HC3N4; (d) EDX mapping of Co/HC3N4, (eh) color images showing the distribution of Co, C, N, and O over the Co/HC3N4 surface.
Figure 2. SEM images of (a) g-C3N4, (b) HC3N4, (c) Co/HC3N4; (d) EDX mapping of Co/HC3N4, (eh) color images showing the distribution of Co, C, N, and O over the Co/HC3N4 surface.
Catalysts 13 01140 g002
Figure 3. XPS analysis of Co/HC3N4 for (a) N 1s, (b) C 1s, (c) Co 2p.
Figure 3. XPS analysis of Co/HC3N4 for (a) N 1s, (b) C 1s, (c) Co 2p.
Catalysts 13 01140 g003
Figure 4. Photocatalytic CO2 reduction with CH4 over g-C3N4, HC3N4, and Co-doped HC3N4 composite samples with a CO2/CH4 ratio of 1.0.
Figure 4. Photocatalytic CO2 reduction with CH4 over g-C3N4, HC3N4, and Co-doped HC3N4 composite samples with a CO2/CH4 ratio of 1.0.
Catalysts 13 01140 g004
Figure 5. Effect of CH4/CO2 feed ratio in a photocatalytic DRM process over a Co/HC3N4 composite on (a) CO evolution and (b) H2 evolution.
Figure 5. Effect of CH4/CO2 feed ratio in a photocatalytic DRM process over a Co/HC3N4 composite on (a) CO evolution and (b) H2 evolution.
Catalysts 13 01140 g005aCatalysts 13 01140 g005b
Figure 6. Photocatalytic CO2 reduction through bi-reforming reactions: (a) CO2 reduction with CH4 and H2O and (b) CO2 reduction with methanol and water.
Figure 6. Photocatalytic CO2 reduction through bi-reforming reactions: (a) CO2 reduction with CH4 and H2O and (b) CO2 reduction with methanol and water.
Catalysts 13 01140 g006
Figure 7. (a) Photocatalytic dry reforming of methane (DRM) over 2% Co/HC3N4 composite per 4 h in consecutive three cycles; (b) XRD analysis of fresh and used 2% Co/HC3N4 composite; (c) FTIR analysis of fresh and used 2% Co/HC3N4 composite.
Figure 7. (a) Photocatalytic dry reforming of methane (DRM) over 2% Co/HC3N4 composite per 4 h in consecutive three cycles; (b) XRD analysis of fresh and used 2% Co/HC3N4 composite; (c) FTIR analysis of fresh and used 2% Co/HC3N4 composite.
Catalysts 13 01140 g007
Figure 8. Proposed mechanism for the photocatalytic dry and bi-reforming of methane to produce CO and H2 over a Co/HC3N4 composite.
Figure 8. Proposed mechanism for the photocatalytic dry and bi-reforming of methane to produce CO and H2 over a Co/HC3N4 composite.
Catalysts 13 01140 g008
Table 1. Performance comparison of different systems producing CO and H2 via a DRM process.
Table 1. Performance comparison of different systems producing CO and H2 via a DRM process.
CatalystFeed RatioProduction Rate
(µmol g−1 h−1) *
Selectivity
(%)
CH4COH2COH2CH4
HC3N4CH4/CO2 = 1.0-30.3623.6956.1743.83
2%Co/HC3N4CH4/CO2 = 1.0-55541.293.096.91
2%Co/HC3N4CH4/CO2: 2:1-67629.795.794.21
2%Co/HC3N4CH4/CO2: 1:2-42023.894.645.36
2%Co/HC3N4CO2-CH4-H2O-1752487.912.1
2%Co/HC3N4CO2-CH3OH-H2O4.4277144463.536.50.4
* Yield rate was calculated based on 4 h of irradiation time and 150 mg of catalyst loading.
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

Tahir, M.; Ali Khan, A.; Bafaqeer, A.; Kumar, N.; Siraj, M.; Fatehmulla, A. Highly Stable Photocatalytic Dry and Bi-Reforming of Methane with the Role of a Hole Scavenger for Syngas Production over a Defective Co-Doped g-C3N4 Nanotexture. Catalysts 2023, 13, 1140. https://doi.org/10.3390/catal13071140

AMA Style

Tahir M, Ali Khan A, Bafaqeer A, Kumar N, Siraj M, Fatehmulla A. Highly Stable Photocatalytic Dry and Bi-Reforming of Methane with the Role of a Hole Scavenger for Syngas Production over a Defective Co-Doped g-C3N4 Nanotexture. Catalysts. 2023; 13(7):1140. https://doi.org/10.3390/catal13071140

Chicago/Turabian Style

Tahir, Muhammad, Azmat Ali Khan, Abdullah Bafaqeer, Naveen Kumar, Mohammad Siraj, and Amanullah Fatehmulla. 2023. "Highly Stable Photocatalytic Dry and Bi-Reforming of Methane with the Role of a Hole Scavenger for Syngas Production over a Defective Co-Doped g-C3N4 Nanotexture" Catalysts 13, no. 7: 1140. https://doi.org/10.3390/catal13071140

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