Next Article in Journal / Special Issue
Mono-, Di-, Tri-Pyrene Substituted Cyclic Triimidazole: A Family of Highly Emissive and RTP Chromophores
Previous Article in Journal
Visible-Light-Active Photocatalysts for Environmental Remediation and Organic Synthesis
Previous Article in Special Issue
Modelling Photoionisations in Tautomeric DNA Nucleobase Derivatives 7H-Adenine and 7H-Guanine: Ultrafast Decay and Photostability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photocatalytic Reduction of CO2 over Iron-Modified g-C3N4 Photocatalysts

1
Centre for Energy and Environmental Technologies (CEET), Institute of Environmental Technology, VSB-Technical University of Ostrava, 17. Listopadu 15/2172, 70800 Ostrava, Czech Republic
2
Institute of Nanoscience and Nanotechnology, NCSR “Demokritos”, Patriarchou Gregoriou E & 27 Neapoleos Str., 15341 Agia Paraskevi, Greece
*
Authors to whom correspondence should be addressed.
Photochem 2021, 1(3), 462-476; https://doi.org/10.3390/photochem1030030
Submission received: 24 October 2021 / Revised: 10 November 2021 / Accepted: 11 November 2021 / Published: 13 November 2021
(This article belongs to the Special Issue Feature Papers in Photochemistry)

Abstract

:
Pure g-C3N4 sample was prepared by thermal treatment of melamine at 520 °C, and iron-modified samples (0.1, 0.3 and 1.1 wt.%) were prepared by mixing g-C3N4 with iron nitrate and calcination at 520 °C. The photocatalytic activity of the prepared materials was investigated based on the photocatalytic reduction of CO2, which was conducted in a homemade batch reactor that had been irradiated from the top using a 365 nm Hg lamp. The photocatalyst with the lowest amount of iron ions exhibited an extraordinary methane and hydrogen evolution in comparison with the pure g-C3N4 and g-C3N4 with higher iron amounts. A higher amount of iron ions was not a beneficial for CO2 photoreduction because the iron ions consumed too many photogenerated electrons and generated hydroxyl radicals, which oxidized organic products from the CO2 reduction. It is clear that there are numerous reactions that occur simultaneously during the photocatalytic process, with several of them competing with CO2 reduction.

Graphical Abstract

1. Introduction

The most pressing problems of the present time are undoubtedly the increasing energy requirements that are closely connected to fossil fuel combustion and the increasing concentration of greenhouse gases in the atmosphere. These problems are related, and finding a new clean source of energy is becoming increasingly important. Developed countries are trying to regulate greenhouse gas emissions; for example, the European Union has a long-term goal to reduce its carbon footprint. The first step towards this goal was to reduce greenhouse gas emissions by 20% (from 1990 levels) by 2020. This goal was successfully achieved, and emissions were reduced by 23% in 2018. However, a new legislation was put in place and added to the Paris Agreement. This legislation aims to reduce greenhouse gas emissions by 40% (from 1990 levels) by 2030. According to Energy in figures—Statistical Pocket Book 2019 issued by European Union, world energy production is increasing every year, and it is predicted that it will continue to rise in the years to come [1]. The highest annual increase of world energy production still comes from solid fossil fuels and natural gas compared to renewable or nuclear sources. Fossil fuel reserves are not infinite, and when considering the rate at which the world’s energy consumption is increasing, it is clear that more focus has to be placed on finding new sources of energy. The photocatalytic reduction of carbon dioxide (CO2) is currently one of the hottest candidates for this purpose. Not only can the products (CO, CH4) of this reaction be used as an energy source, but the reaction utilizes CO2, which is the most abundant source of carbon and is also considered to be a waste; therefore, it is not utilized [2]. The photocatalytic community has been studying the photocatalytic reduction of CO2 for years. While the exact mechanism is not known, many important advances have been made [3,4].
There have been many photocatalysts that have been discovered, and different combinations of these photocatalysts have been tested to determine which combination produces the best photocatalytic reduction of CO2 over the years [5,6]; however, most attention has focused on TiO2 [7,8,9]. TiO2 has quite large band gap energy, and more importantly, the conduction band potential is not in a convenient position located at around −0.5 V (vs. NHE at pH = 7), which is very close to the required redox potentials of essential reactions such as CO2/HCOOH (−0.61 V) CO2/HCHO (−0.48 V), 2H+/H2 (−0.41 V), and CO2/CH4 (−0.24 V) [7]. Finding a new photocatalyst is an immediate goal that most of the photocatalytic community is focused on.
Graphitic carbon nitride (g-C3N4) has attracted increasing attention from the photocatalytic community, and research groups dealing with the photocatalytic reduction of CO2 have already noticed its potential [10,11,12]. Its indirect band gap energy (around 2.7 eV) is especially narrow, and it has a conduction band potential (around −1.4 V) that makes it promising [13,14]. g-C3N4 has also high chemical stability, and more importantly, it has much more suitable potential of valence and conduction bands than TiO2 [15]. The photocatalytic community is trying to shift the reaction into the visible light region, and the band gap energy of 2.7 eV can allow that. This corresponds approximately to a wavelength of 460 nm, which is somewhere on the edge between the blue and green region of visible light.
Even though g-C3N4 is a promising photocatalyst, it has a lot of drawbacks, such as low specific surface area, the fast recombination rate of generated charge carriers, and the particle boundary effects that interrupt the delocalization of electrons [16]. The advantages and disadvantages of g-C3N4 are summarized and discussed in the following reviews [16,17].
Many research teams have focused their attention to overcome the disadvantages of g-C3N4. Various modifications of g-C3N4 have been investigated in order to increase photocatalytic activity. For example, the specific surface area of g-C3N4 can be significantly enhanced by exfoliation [18,19]. There are several possibilities and techniques that can be used to achieve the exfoliation [20,21]. Another method for enhancing the photocatalytic activity of g-C3N4 is metal modification, which is mainly used as an effective approach to tune the electrical and optical properties [22]. Noble metals are the first and obvious choice [10,23]. However, noble metals are expensive, and therefore, other metal modification possibilities have been investigated. One such possibility is zero-valent iron, which is a cost-effective alternative, and g-C3N4 with incorporated iron particles has shown the improved absorption of visible light and as well as the lower recombination of charge carriers [24]. Modification caused by iron can also create a synergetic system that combines the Fenton and photocatalytic processes. When iron is present in a trivalent form, Fenton process can proceed [25]. Since there are a lot of lone electron pairs in the g-C3N4 heptazine structure, bonding with a foreign metal cation is relatively easy [26].
Although g-C3N4 only attracted the attention of the photocatalytic community just a few years ago, numerous papers about various modifications of g-C3N4 for the photocatalytic reduction of CO2 can be found [11,27]. However, to the best of our knowledge, there is only a very limited number of works describing the utilization of g-C3N4/Fe for this application. In the present work, iron-modified g-C3N4 photocatalysts were synthesized by a simple in situ method. The photocatalysts were characterized and tested for photocatalytic CO2 reduction where selectivity towards certain products was revealed and discussed.

2. Materials and Methods

2.1. Preparation

Melamine powder 99% was purchased from Alfa Aesar, and iron nitrate 98% was purchased from Chem-Lab.
Bulk g-C3N4 was synthesized by thermal polycondensation of melamine. The conditions were chosen by taking the relevant literature into account [28,29,30,31], and the results of our previous work were also considered [32,33]. Typically, 5 g of melamine was put into a covered alumina crucible and was heated in a muffle furnace in air to 520 °C with a rate of 5 °C/min, where it remained for 2 h. The yellow product (yield~2.4 g) was collected and was ground into fine powder. The sample was called CN.
A similar procedure was used for the preparation of the iron-modified g-C3N4 photocatalysts. In detail, 5 g of melamine was mixed with specific amounts of iron nitrate resulting in iron concentrations of 0.1, 0.3, and 1.1 wt.%. The mixture was stirred in 100 mL H2O for 1 h to achieve a homogeneous suspension. After drying at 80 °C in an oven, the powder was put into a covered alumina crucible and was heated in a muffle furnace in air to 520 °C at a rate of 5 °C/min, where it remained for 2 h. The product (yield~2.5 g) was collected and ground into fine powder. The samples were called CN-Fex, where x is the amount (wt.%) of iron.

2.2. Characterization

The crystalline structure, surface area, porosity, chemical composition, and optical characteristics of the samples were thoroughly examined, and the nature of the iron ions was also investigated. Details about the characterization techniques and the equipment that was used are presented in the Supplementary Materials.

2.3. Photocatalytic Test

The photocatalytic reduction of CO2 was performed in a homemade batch-stirred stainless steel reactor (volume 348 mL) with a quartz glass window on the top (Figure S1). A UV 8 W Hg lamp (peak intensity at 365 nm; Ultra-Violet Products Inc., Cambridge, UK) was used as a light source, and it was located over the quartz glass window of the reactor at a specific height so that the intensity on the level of the suspension was 0.833 mW/cm2.
The reactor was filled with 0.09 g of each photocatalyst and 100 mL of 0.2 M NaOH. The NaOH was added in order to increase the solubility of CO2 in water, thus facilitating the photocatalytic reaction. The suspension was stirred with a magnetic stirrer the entire time to prevent the sedimentation of the photocatalyst. Before the photocatalytic reaction began, the reactor was tightly closed and purged with (He or) CO2. The pH of the suspension decreased during the purging from ~12.5 to ~6.9. A detailed description of the experimental procedure is available in the Supplementary Materials.
The gaseous samples were analyzed in 2 h intervals between the duration 0–8 h, where 0 corresponded to the moment prior to UV irradiation. Samples were collected using a gastight syringe (Hamilton Co., Reno, NV, USA) and were analyzed on a gas chromatograph (Shimadzu Tracera GC-2010 Plus) equipped with a BID detector. Each sample was measured at least three times using the same batch (same photocatalyst, same NaOH) in order to confirm the durability of the photocatalyst and the reproducibility of the experiments (within 5% error). The photon flux and apparent quantum yields (AQY) for individual products in the presence of each photocatalyst were calculated. The photocatalysts were first examined in an inert atmosphere to confirm that no products were formed, and after that the photocatalytic reduction of CO2 was conducted.

3. Results and Discussion

3.1. Crystalline Structure

The XRD patterns of the g-C3N4 photocatalysts are presented in Figure 1. For all of the samples, the two characteristic diffraction peaks of g-C3N4 are observed (JCPDS, PDF #87-1526). The weak peak at 13.1° corresponds to the (100) plane, which is related to the in-plane structural packing motif of the tri-s-triazine units with an interplanar distance of d = 0.675 nm [34,35]. The strong peak at 27.5° corresponds to the (002) plane, which is attributed to the interlayer stacking of aromatic rings with a distance of d = 0.324 nm [36,37]. After the modification, no additional peaks corresponding to iron are observed, which was probably due to the low initial amount of iron nitrate. Furthermore, there are no significant changes to the position or the intensity of the peaks, indicating that the crystalline structure of the materials is preserved [38].

3.2. Surface Area and Porosity

The N2 adsorption–desorption isotherms and pore-size distributions of the g-C3N4 photocatalysts are displayed in Figure S2, with the characteristic values given in Table 1. All of the samples showed characteristic hysteresis loops of type IV isotherms, which is typical for mesoporous materials [39]. After the modification, the specific surface area of the photocatalysts decreased, which can be attributed to the partial blocking of the g-C3N4 pores by the iron. On the other hand, the pore size distribution remained practically unchanged across all of the samples.

3.3. Chemical Composition

The FT-IR analysis results are shown in Figure 2. For all of the samples, the known characteristic peaks of g-C3N4 can be observed. The broad peak at the 3500–3000 cm−1 region corresponds to the stretching mode of all of the OH-containing species and N-H stretching from the residual amino groups [40,41]. The shoulder located at 1630 cm−1 represents the H-O-H bending mode of all of the water molecules present at the surface of the material. The peaks in the 1628–1232 cm−1 region correspond to the characteristic stretching of the C-N heterocycles, including the trigonal N-(C)3 and bridging H-N-(C)2 units [42]. Finally, the sharp absorption peak at 805 cm−1 can be attributed to the breathing mode of the triazine units [43]. The FT-IR spectra of the modified materials demonstrate no significant changes. More specifically, there is no observable shift in the peaks, indicating that the chemical bonding of the main g-C3N4 network is unaffected [38].

3.4. Optical Characteristics

The measured diffuse reflectance spectra of the g-C3N4 photocatalysts are shown in the inset of Figure 3. They were used for the construction of the absorption functions (F × E)1/2 = f(E) presented in Figure 4, which allowed the determination of the materials’ band gap energy (Eg), as described in Ref. [44]. It is evident that after modification, the light absorption of the materials is increased. However, the Eg only slightly decreased from 2.73 eV for bulk g-C3N4, which is in agreement with the literature [16,17], to 2.71 eV for the modified samples.
The properties of the photogenerated charge carriers were evaluated with fluorescence measurements at λex = 365 nm (Figure 4). All of the samples showed a broad band with λem at around 455 nm, which is in agreement with the absorption edge wavelength measured by UV-vis spectroscopy. Such a signal can be attributed to the band–band PL phenomenon, which mainly results from the n-π* electronic transitions in g-C3N4 [45]. The strongest emission intensity is displayed by bulk g-C3N4, which suggests a faster electron–hole radiative recombination rate. On the other hand, the modified samples show significantly weaker emission intensity, which becomes even more prominent as the iron content increases. This indicates a lower recombination rate and higher charge transfer efficiency, suggesting a potential improvement in the photocatalytic activity of the materials.
Photoelectrochemical measurements are one of the characterization techniques that allow the prediction of the photocatalyst behavior. Since the photocatalyst is irradiated under an external potential that has been applied to the working electrode, the recombination of electrons and holes is strongly suppressed. Based on the photocurrent generation (Figure 5), it is clear that the highest amount of charge carriers is produced in the case of bulk g-C3N4. This is connected to a higher specific surface area. However, a higher amount of produced charge carriers does not mean higher photocatalytic activity, especially since bulk g-C3N4 has the highest recombination rate of charge carriers (Figure 4). It is known that the redox potential of Fe3+/Fe2+ is below the CB of g-C3N4. After iron modification, the photogenerated electrons could be trapped by the Fe2+ sites, leading to the reduced recombination of the photogenerated electron–hole pairs. With a higher iron concentration, more photogenerated electron trapping sites exist, thus leading to the further decrease of the PL intensity.

3.5. Mössbauer and EPR Spectroscopy

Mössbauer spectroscopy was used to investigate the nature of the iron of the modified g-C3N4 photocatalysts. The 80 K spectra of samples CN-Fe0.3 and CN-Fe1.1 are shown in Figure 6. The most prominent characteristic of both spectra are the two distinct doublets that are characteristic of the Fe2+ ions in two different chemical environments [46]. These environments may be associated with the iron that is present in the -C3N4 network gaps or on the layer surface. The spectra also show a lower intensity doublet showing the presence of Fe3+ ions. A subspectrum that can be attributed to magnetic Fe2O3 [47], which is most likely due to the reaction of weakly stabilized Fe3+/2+ ions with air, is also observed in all of the spectra. At 80 K there is no sign of any subspectral broadening that would suggest the existence of superparamagnetic nanoparticles. Thus, the only crystalline iron-containing phase observed through Mössbauer spectroscopy is that of the Fe2O3 oxide.
By comparing the relative spectral areas of the corresponding peaks (Table S2), it is evident that iron in the g-C3N4 matrix is mainly present in the form of Fe2+ species. This indicates that during melamine polycondensation, the Fe3+ ions of the iron nitrate precursor are reduced to Fe2+. It should be also noted that the Mössbauer peaks of the sample CN-Fe0.3 are less intense than those of sample CN-Fe1.1, which confirms the lower amount of Fe present in the photocatalyst. The EPR spectra of the modified samples (Figure S3) display a broad line that is characteristic of iron ion–ion interactions [48]. However, as expected, the Fe2+ ions were not visible in the EPR measurements.
Recent studies have revealed that transition metal ions such as Au3+, Ag+, Cu2+, and Fe3+ may be reduced during g-C3N4 synthesis at ~550 °C [49,50]. During melamine polycondensation, apart from NH3, reactive species such as CNH2, H2NCN, and CN2+ are also released. These species can progressively reduce Fe3+ to Fe2+ until the formed ions are stabilized in the g-C3N4 network [51].

3.6. Photocatalytic Activity

Figure 7 shows the dependence of product yields on the irradiation time (0–8 h). The main products of the photocatalytic reduction of CO2 are methane (Figure 7a) and carbon monoxide (Figure 7b). However, the hydrogen generated from the photocatalytic water splitting is also present in much higher concentrations (Figure 7c).
Photocatalytic CO2 reduction is a complex process with a number of reactions occurring simultaneously. The g-C3N4 band gap energy around 2.7 eV (Figure 3) corresponds to a wavelength of approximately 460 nm, and since the photocatalytic reduction was conducted under the irradiation of 365 nm, each incident photon should lead to the generation of an electron and hole. The product yields and the apparent quantum yields (AQY) are shown in the Supplementary Materials (Table S1). Among the photocatalysts, the CN-Fe0.1 demonstrated the highest CH4 and H2 AQY at 0.0020% and 0.0733%, respectively. One of the most important properties of a photocatalyst is the CB and VB edge potentials. XPS analysis was used to determine the potential of the valence bands of each photocatalyst (Figure 8).
As can be seen, the VB edge potential shifted to more positive values after the modification, which indicates the increased oxidation efficiency of the photogenerated holes [21]. By combining the band gap energy and VB edge potential values, the CB edge potentials of the materials were also calculated. Thus, the electronic band structures of the g-C3N4 photocatalysts were created (Figure 9). As the Eg of the photocatalysts remains practically unchanged after the modification, the CB edge potential shifts to less negative values, indicating a slight decrease in the reduction strength of the photogenerated electrons. However, the potential of CB is still higher than the redox potential that is required for the CO2 reactions (Equations (1)–(5)).
The potentials of valence and conduction bands slightly differ for each g-C3N4 material; however, it is approximately inside the interval −1.5 V (CB potential) and 1.2 V (VB potential) vs. NHE (pH = 7). It is the relatively high reduction strength (negative potential of CB) that makes g-C3N4 such an attractive material.
The potential of CB is negative enough to allow any of the partial redox reactions to proceed (Equations (1)–(5)) [52,53]. However, the direct reduction of a CO2 molecule by a single electron is not possible due to a very negative required potential of −1.9 V (Equation (6)) [2,5,54,55].
C O 2 + 2 e + 2 H + H C O O H   E 0 = 0.61   V
C O 2 + 2 e + 2 H + C O + H 2 O   E 0 = 0.53   V
C O 2 + 4 e + 4 H + H C H O + H 2 O   E 0 = 0.48   V
C O 2 + 6 e + 6 H + C H 3 O H + H 2 O   E 0 = 0.38   V  
C O 2 + 8 e + 8 H + C H 4 + 2 H 2 O   E 0 = 0.24   V
C O 2 + e C O 2   E 0 = 1.90   V
All of the redox potentials are stated vs. NHE at pH = 7.
Based on the equations above, it is clear that the photocatalytic reduction of CO2 is a multielectron process that also requires the presence of a hydrogen cation. The necessity of a hydrogen cation is the reason why the reaction has to be conducted in aqueous phase or in the presence of water vapor.
There are two possible ways that water can be oxidized by photogenerated holes toward H+ (Equations (7) and (8)) [56].
2 H 2 O + 4 h + O 2 + 4 H +   E 0 = + 0.82   V
H 2 O + h +   ° O H + H +   E 0 = + 2.32   V
It is clear that Equation (8) would be more probable since it requires just one hole and one water molecule; however, its redox potential is much more positive than the VB potential of g-C3N4, and therefore, this reaction cannot proceed in the presence of g-C3N4 material, leaving Equation (7) as the main source of the required H+ ions. This significantly complicates the situation. First of all, the reaction (Equation (7)) is much less probable since it requires two water molecules o and four holes at the same time, but oxygen is also produced along with H+. The reduction of oxygen molecules to a superoxide radical (Equation (9)) is one of the competitive reactions for the reduction of CO2 [27,57].
O 2 + e   ° O 2   E 0 = 0.33   V
The presence of oxygen not only competes with CO2 molecules to be adsorbed on the surface of the photocatalyst but also consumes the necessary electrons that are needed for the reduction.
Another competitive reaction is the reduction of the generated H+ ions to hydrogen (Equation (10)).
2 H + + 2 e H 2   E 0 = 0.41   V
It is this competitive reaction that is responsible for the presence of hydrogen among the detected products (Figure 7c). Methane and carbon monoxide were other detected reaction products that can be generated according to Equations (2) and (5). Unfortunately, the possible products in the liquid phase were below the detection limit, and therefore, their presence could not be confirmed.
It is clear that the photocatalytic reduction of CO2 is a very complex process that follows several multielectron steps. Nevertheless, the presence of the Fe3+/Fe2+ couple (Figure 6) can complicate the situation even more. Many publications say that the presence of iron ions can be beneficial due to the possibility of photo-Fenton process occurring along with the photocatalysis [58,59,60]. What does this mean in the case of a photocatalytic CO2 reduction? Part of the photogenerated electrons can directly transfer to Fe3+ ions and form Fe2+ (Equation (11)).
F e 3 + + e F e 2 +   E 0 = 0.77   V
Holes, however, stay in the valence band and can participate in the oxidation reaction. The presence of Fe3+ can improve the separation of generated charge carriers; however, the negative side effect is the consumption of electrons needed for the photocatalytic reduction of CO2; therefore, finding an optimum amount of iron ions is necessary [15].
Since the Fe2+ ions are unstable, they easily oxidize back into Fe3+ in the presence of oxygen and form either O 2 (Equation (12)) or H2O2 (Equation (13)).
F e 2 + + O 2 + e F e 3 + + O 2 °   E 0 = 0.05   V
2 F e 2 + + O 2 + 2 H + 2 F e 3 + + H 2 O 2   E 0 = + 0.68   V
The above-mentioned equations clearly suggest the consumption of electrons in both redox Fe3+/Fe2+ reactions. Higher amounts of iron ions would clearly explain the lower yields of the CO2 reduction products. On the other hand, the oxygen used in Equations (12) and (13) does not compete with CO2 molecules needed for adsorption on the surface of the photocatalyst.
In addition, the holes in g-C3N4 VB have sufficient potential to oxidize H2O to H2O2 (+0.69 V vs. NHE) [27,61] and the potential of H2O2/·OH (1.07 V vs. NHE) [27] is more positive than the redox potential of Fe3+/Fe2+ (0.77 V vs. NHE) [27,61,62]; the produced H2O2 reacts with Fe2+ to produce ·OH species. Hydroxyl radical species are very strong oxidizing agents and are very beneficial when the removal of organics in water is the goal. However, the goal is to create organic compounds for the photocatalytic reduction of CO2. Therefore, a higher amount of Fe would lead to a higher amount of hydroxyl radicals that would be able to oxidize the already generated organic products (CH4) [27]. This is the reason why the CH4 yields were lower in the case of photocatalysts containing 0.3 and 1.1 wt.% of Fe (Table S1). Furthermore, when the iron content exceeds an optimal value, the photocatalytic performance decreases, which is possibly due to the competitive capture between the adsorbed CO2 and Fe2+ sites. In this case, an excess of Fe2+ sites could trap more photogenerated electrons, leading to fewer electrons being available to react with the adsorbed CO2 molecules. On the other hand, the higher Fe content in the photocatalyst clearly leads to a higher selectivity toward CO production (Figure 7b). This suggests higher selectivity for CO2 reduction to CO. Higher amounts of iron ions are incorporated into the g-C3N4 lattice and affect its electronic properties, which were confirmed by the valence band position shift after the addition of Fe (Figure 9), forming impurities that lead to an enhanced separation of charge carriers. On the other hand, this means a higher amount of hydroxyl radicals and a higher rate of reverse oxidation for the produced hydrocarbons [63].

4. Conclusions

Iron-modified g-C3N4 photocatalysts were prepared using a simple thermal treatment of melamine mixed with specific amounts of iron nitrate. The XRD and FT-IR measurements did not reveal any iron in the photocatalysts due to its very low concentrations (0.1, 0.3 and 1.1 wt.%). However, its presence in the form of Fe2+, Fe3+, and Fe2O3 was demonstrated by Mössbauer spectroscopy, with Fe2+ being the main species.
The addition of iron resulted in a decrease in the photocurrent and specific surface area due to the partial blocking of the g-C3N4 pores by iron. However, this led to the reduced recombination of the photogenerated electron–hole pairs, which is important for high photocatalytic efficiency.
The photocatalytic activity measurements of the prepared materials in the CO2 reduction showed that the photocatalyst with the lowest amount of iron exhibited the biggest methane and hydrogen evolution in comparison with pure g-C3N4 and g-C3N4 with higher iron amounts. Excess iron consumes electrons that are needed for the reduction reactions and generates hydroxyl radicals that oxidize organic products from the CO2 reduction. The iron-modified (0.1 wt.%) g-C3N4 proved to be a promising photocatalyst for CO2 conversion into valuable hydrocarbons.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/photochem1030030/s1, Figure S1: The batch reactor used for the photocatalytic reduction of CO2, Figure S2: N2 adsorption-desorption isotherms (a) and pore size distribution curves (b) of the g-C3N4 photocatalysts, Figure S3: EPR spectra of samples CN-Fe0.3 and CN-Fe1.1, Table S1: Recalculated product yields to μmol·s−1 and apparent quantum yields (AQY) for each product in the presence of the photocatalysts, Table S2: Mössbauer parameters for samples CN-Fe0.3 and CN-Fe1.1 and corresponding assignments. Isomer Shift (IS), line width (HWHM), Quadrupole Splitting (QS), Hyperfine Magnetic Feld (HMF) and the subspectral area (= relative amount of iron).

Author Contributions

Conceptualization, I.P., C.T., and K.K.; methodology, I.P.; software, M.E.; validation, K.K. and M.R.; formal analysis, M.R.; investigation, N.I., E.D., and P.D.; resources, C.T. and K.K.; data curation, N.T. and T.G.; writing—original draft preparation, M.E., M.R.; writing—review and editing, M.E., M.R., and I.P.; visualization, I.P.; supervision, C.T. and K.K.; project administration, C.T. and K.K.; funding acquisition, C.T. and K.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was co-financed by the European Union and Greek national funds through the Operational Program Competitiveness, Entrepreneurship and Innovation, under the call RESEARCH-CREATE-INNOVATE (project code: T1EDK-05545) and by EU structural funding in Operational Programme Research Development and Education, project No. CZ.02.1.01/0.0/0.0/16_019/0000853 “IET-ER” and by using Large Research Infrastructure ENREGAT supported by the Ministry of Education, Youth and Sports of the Czech Republic under project No. LM2018098. The authors would also like to acknowledge the Hellenic Foundation for Research and Innovation and the General Secretariat for Research and Innovation for Grant number 1468.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article.

Acknowledgments

The authors would like to thank Andreas Karydas and Kalliopi Tsampa from “XRF Laboratory” of Institute of Nuclear and Particle Physics at NCSR “Demokritos” for the XRF measurements.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. European Commission. EU Energy in Figures-Statistical Pocket Book 2019; Drukarnia INTERAK SP ZOO: Czarnków, Poland, 2019. [Google Scholar]
  2. Vu, N.N.; Kaliaguine, S.; Do, T.O. Critical Aspects and Recent Advances in Structural Engineering of Photocatalysts for Sunlight-Driven Photocatalytic Reduction of CO2 into Fuels. Adv. Funct. Mater. 2019, 29, 1901825. [Google Scholar] [CrossRef]
  3. Alkhatib, I.I.; Garlisi, C.; Pagliaro, M.; Al-Ali, K.; Palmisano, G. Metal-organic frameworks for photocatalytic CO2 reduction under visible radiation: A review of strategies and applications. Catal. Today 2020, 340, 209–224. [Google Scholar] [CrossRef]
  4. Yuan, L.; Xu, Y.J. Photocatalytic conversion of CO2 into value-added and renewable fuels. Appl. Surf. Sci. 2015, 342, 154–167. [Google Scholar] [CrossRef]
  5. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic reduction of CO2 on TiO2 and other semiconductors. Angew. Chem. 2013, 52, 7372–7408. [Google Scholar] [CrossRef] [PubMed]
  6. Nikokavoura, A.; Trapalis, C. Alternative photocatalysts to TiO2 for the photocatalytic reduction of CO2. Appl. Surf. Sci. 2017, 391, 149–174. [Google Scholar] [CrossRef]
  7. 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]
  8. Low, J.; Cheng, B.; Yu, J. Surface modification and enhanced photocatalytic CO2 reduction performance of TiO2: A review. Appl. Surf. Sci. 2017, 392, 658–686. [Google Scholar] [CrossRef]
  9. Akhter, P.; Farkhondehfal, M.A.; Hernández, S.; Hussain, M.; Fina, A.; Saracco, G.; Khan, A.U.; Russo, N. Environmental issues regarding CO2 and recent strategies for alternative fuels through photocatalytic reduction with titania-based materials. J. Environ. Chem. Eng. 2016, 4, 3934–3953. [Google Scholar] [CrossRef]
  10. Kočí, K.; Van, H.D.; Edelmannová, M.; Reli, M.; Wu, J.C.S. Photocatalytic reduction of CO2 using Pt/C3N4 photocatalyts. Appl. Surf. Sci. 2020, 503, 144426. [Google Scholar] [CrossRef]
  11. Li, X.; Xiong, J.; Gao, X.; Huang, J.; Feng, Z.; Chen, Z.; Zhu, Y. Recent advances in 3D g-C3N4 composite photocatalysts for photocatalytic water splitting, degradation of pollutants and CO2 reduction. J. Alloys Compd. 2019, 802, 196–209. [Google Scholar] [CrossRef]
  12. Xiang, Q.; Cheng, B.; Yu, J. Graphene-Based Photocatalysts for Solar-Fuel Generation. Angew. Chem. Int. Ed. Engl. 2015, 54, 11350–11366. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, J. Effect of phosphorus doping on electronic structure and photocatalytic performance of g-C3N4: Insights from hybrid density functional calculation. J. Alloys Compd. 2016, 672, 271–276. [Google Scholar] [CrossRef]
  14. Wu, H.Z.; Liu, L.M.; Zhao, S.J. The effect of water on the structural, electronic and photocatalytic properties of graphitic carbon nitride. Phys. Chem. Chem. Phys. 2014, 16, 3299–3304. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, Q.; Guo, Y.; Chen, Z.; Zhang, Z.; Fang, X. Constructing a novel ternary Fe(III)/graphene/g-C3N4 composite photocatalyst with enhanced visible-light driven photocatalytic activity via interfacial charge transfer effect. Appl. Catal. B 2016, 183, 231–241. [Google Scholar] [CrossRef]
  16. Prasad, C.; Tang, H.; Bahadur, I. Graphitic carbon nitride based ternary nanocomposites: From synthesis to their applications in photocatalysis: A recent review. J. Mol. Liq. 2019, 281, 634–654. [Google Scholar] [CrossRef]
  17. Xiao, M.; Luo, B.; Wang, S.; Wang, L. Solar energy conversion on g-C3N4 photocatalyst: Light harvesting, charge separation, and surface kinetics. J. Energy Chem. 2018, 27, 1111–1123. [Google Scholar] [CrossRef] [Green Version]
  18. Praus, P.; Svoboda, L.; Dvorský, R.; Reli, M.; Kormunda, M.; Mančík, P. Synthesis and properties of nanocomposites of WO3 and exfoliated g-C3N4. Ceram. Int. 2017, 43, 13581–13591. [Google Scholar] [CrossRef]
  19. Todorova, N.; Papailias, I.; Giannakopoulou, T.; Ioannidis, N.; Boukos, N.; Dallas, P.; Edelmannová, M.; Reli, M.; Kočí, K.; Trapalis, C. Photocatalytic H2 Evolution, CO2 Reduction, and NOx Oxidation by Highly Exfoliated g-C3N4. Catalysts 2020, 10, 1147. [Google Scholar] [CrossRef]
  20. Mohamed, N.A.; Safaei, J.; Ismail, A.F.; Noh, M.F.M.; Arzaee, N.A.; Mansor, N.N.; Ibrahim, M.A.; Ludin, N.A.; Sagu, J.S.; Teridi, M.A.M. Fabrication of exfoliated graphitic carbon nitride, (g-C3N4) thin film by methanolic dispersion. J. Alloys Compd. 2020, 818, 152916. [Google Scholar] [CrossRef]
  21. Papailias, I.; Todorova, N.; Giannakopoulou, T.; Ioannidis, N.; Boukos, N.; Athanasekou, C.P.; Dimotikali, D.; Trapalis, C. Chemical vs thermal exfoliation of g-C3N4 for NOx removal under visible light irradiation. Appl. Catal. B 2018, 239, 16–26. [Google Scholar] [CrossRef]
  22. Wen, J.; Xie, J.; Chen, X.; Li, X. A review on g-C3N4-based photocatalysts. Appl. Surf. Sci. 2017, 391, 72–123. [Google Scholar] [CrossRef]
  23. Masih, D.; Ma, Y.; Rohani, S. Graphitic C3N4 based noble-metal-free photocatalyst systems: A review. Appl. Catal. B 2017, 206, 556–588. [Google Scholar] [CrossRef]
  24. Heidarpour, H.; Padervand, M.; Soltanieh, M.; Vossoughi, M. Enhanced decolorization of rhodamine B solution through simultaneous photocatalysis and persulfate activation over Fe/C3N4 photocatalyst. Chem. Eng. Res. Des. 2020, 153, 709–720. [Google Scholar] [CrossRef]
  25. Hu, J.; Zhang, P.; Cui, J.; An, W.; Liu, L.; Liang, Y.; Yang, Q.; Yang, H.; Cui, W. High-efficiency removal of phenol and coking wastewater via photocatalysis-Fenton synergy over a Fe-g-C3N4 graphene hydrogel 3D structure. J. Ind. Eng. Chem. 2020, 84, 305–314. [Google Scholar] [CrossRef]
  26. Ma, T.; Shen, Q.; Xue, B.Z.J.; Guan, R.; Liu, X.; Jia, H.; Xu, B. Facile synthesis of Fe-doped g-C3N4 for enhanced visible-light photocatalytic activity. Inorg. Chem. Commun. 2019, 107, 107451. [Google Scholar] [CrossRef]
  27. Sun, Z.; Wang, H.; Wu, Z.; Wang, L. g-C3N4 based composite photocatalysts for photocatalytic CO2 reduction. Catal. Today 2018, 300, 160–172. [Google Scholar] [CrossRef]
  28. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef]
  29. Zhang, Y.; Pan, Q.; Chai, G.; Liang, M.; Dong, G.; Zhang, Q.; Qiu, J. Synthesis and luminescence mechanism of multicolor-emitting g-C3N4 nanopowders by low temperature thermal condensation of melamine. Sci. Rep. 2013, 3, 1943. [Google Scholar] [CrossRef]
  30. Ge, L. Synthesis and photocatalytic performance of novel metal-free g-C3N4 photocatalysts. Mater. Lett. 2011, 65, 2652–2654. [Google Scholar] [CrossRef]
  31. Baudys, M.; Paušová, Š.; Praus, P.; Brezová, V.; Dvoranová, D.; Barbieriková, Z.; Krýsa, J. Graphitic Carbon Nitride for Photocatalytic Air Treatment. Materials 2020, 13, 3038. [Google Scholar] [CrossRef] [PubMed]
  32. Papailias, I.; Giannakopoulou, T.; Todorova, N.; Demotikali, D.; Vaimakis, T.; Trapalis, C. Effect of processing temperature on structure and photocatalytic properties of g-C3N4. Appl. Surf. Sci. 2015, 358, 278–286. [Google Scholar] [CrossRef]
  33. Praus, P.; Svoboda, L.; Ritz, M.; Troppová, I.; Šihor, M.; Kočí, K. Graphitic carbon nitride: Synthesis, characterization and photocatalytic decomposition of nitrous oxide. Mater. Chem. Phys. 2017, 193, 438–446. [Google Scholar] [CrossRef]
  34. Yan, H.; Chen, Y.; Xu, S. Synthesis of graphitic carbon nitride by directly heating sulfuric acid treated melamine for enhanced photocatalytic H2 production from water under visible light. Int. J. Hydrogen Energy 2012, 37, 125–133. [Google Scholar] [CrossRef]
  35. Ho, W.; Zhang, Z.; Xu, M.; Zhang, X.; Wang, X.; Huang, Y. Enhanced visible-light-driven photocatalytic removal of NO: Effect on layer distortion on g-C3N4 by H2 heating. Appl. Catal. B 2015, 179, 106–112. [Google Scholar] [CrossRef]
  36. Zhao, H.; Chen, X.; Jia, C.; Zhou, T.; Qu, X.; Jian, J.; Xu, Y.; Zhou, T. A facile mechanochemical way to prepare g-C3N4. Mater. Sci. Eng. B 2005, 122, 90–93. [Google Scholar] [CrossRef]
  37. Zhao, H.; Yu, H.; Quan, X.; Chen, S.; Zhang, Y.; Zhao, H.; Wang, H. Fabrication of atomic single layer graphitic-C3N4 and its high performance of photocatalytic disinfection under visible light irradiation. Appl. Catal. B 2014, 153, 46–50. [Google Scholar] [CrossRef]
  38. Papailias, I.; Todorova, N.; Giannakopoulou, T.; Ioannidis, N.; Dallas, P.; Dimotikali, D.; Trapalis, C. Novel torus shaped g-C3N4 photocatalysts. Appl. Catal. B 2020, 268, 118733. [Google Scholar] [CrossRef]
  39. Papailias, I.; Todorova, N.; Giannakopoulou, T.; Karapati, S.; Boukos, N.; Dimotikali, D.; Trapalis, C. Enhanced NO2 abatement by alkaline-earth modified g-C3N4 nanocomposites for efficient air purification. Appl. Surf. Sci. 2018, 430, 225–233. [Google Scholar] [CrossRef]
  40. Zhao, Z.; Sun, Y.; Luo, Q.; Dong, F.; Li, H.; Ho, W.K. Mass-Controlled Direct Synthesis of Graphene-like Carbon Nitride Nanosheets with Exceptional High Visible Light Activity. Less is Better. Sci. Rep. 2015, 5, 14643. [Google Scholar] [CrossRef]
  41. Zhu, B.; Xia, P.; Ho, W.; Yu, J. Isoelectric point and adsorption activity of porous g-C3N4. Appl. Surf. Sci. 2015, 344, 188–195. [Google Scholar] [CrossRef]
  42. Cao, Y.; Li, Q.; Wang, W. Construction of a crossed-layer-structure MoS2/g-C3N4 heterojunction with enhanced photocatalytic performance. RSC Adv. 2017, 7, 6131–6139. [Google Scholar] [CrossRef] [Green Version]
  43. Li, Y.; Zhang, H.; Liu, P.; Wang, D.; Li, Y.; Zhao, H. Cross-Linked g-C3N4/rGO Nanocomposites with Tunable Band Structure and Enhanced Visible Light Photocatalytic Activity. Small 2013, 9, 3336–3344. [Google Scholar] [CrossRef] [Green Version]
  44. Giannakopoulou, T.; Todorova, N.; Romanos, G.; Vaimakis, T.; Dillert, R.; Bahnemann, D.; Trapalis, C. Composite hydroxyapatite/TiO2 materials for photocatalytic oxidation of NOx. Mater. Sci. Eng. B 2012, 177, 1046–1052. [Google Scholar] [CrossRef]
  45. Hu, S.; Jin, R.; Lu, G.; Liu, D.; Gui, J. The properties and photocatalytic performance comparison of Fe3+-doped g-C3N4 and Fe2O3/g-C3N4 composite catalysts. RSC Adv. 2014, 4, 24863–24869. [Google Scholar] [CrossRef]
  46. Oliveira, A.A.S.; Teixeira, I.F.; Ribeiro, L.P.; Lorençon, E.; Ardisson, J.D.; Fernandez-Outon, L.; Macedo, W.A.A.; Moura, F.C.C. Magnetic amphiphilic nanocomposites produced via chemical vapor deposition of CH4 on Fe-Mo/nano-Al2O3. Appl. Catal. A 2013, 456, 126–134. [Google Scholar] [CrossRef]
  47. Cunha, I.T.; Teixeira, I.F.; Albuquerque, A.S.; Ardisson, J.D.; Macedo, W.A.A.; Oliveira, H.S.; Tristão, J.C.; Sapag, K.; Lago, R.M. Catalytic oxidation of aqueous sulfide in the presence of ferrites (MFe2O4, M=Fe, Cu, Co). Catal. Today 2016, 259, 222–227. [Google Scholar] [CrossRef]
  48. Cuello, N.I.; Elías, V.R.; Winkler, E.; Pozo-López, G.; Oliva, M.I.; Eimer, G.A. Magnetic behavior of iron-modified MCM-41 correlated with clustering processes from the wet impregnation method. J. Magn. Magn. Mater. 2016, 407, 299–307. [Google Scholar] [CrossRef]
  49. Wang, N.; Han, Z.; Fan, H.; Ai, S. Copper nanoparticles modified graphitic carbon nitride nanosheets as a peroxidase mimetic for glucose detection. RSC Adv. 2015, 5, 91302–91307. [Google Scholar] [CrossRef]
  50. Li, Z.; Kong, C.; Lu, G. Visible Photocatalytic Water Splitting and Photocatalytic Two-Electron Oxygen Formation over Cu- and Fe-Doped g-C3N4. J. Phys. Chem. C 2016, 120, 56–63. [Google Scholar] [CrossRef]
  51. Bicalho, H.A.; Lopez, J.L.; Binatti, I.; Batista, P.F.R.; Ardisson, J.D.; Resende, R.R.; Lorençon, E. Facile synthesis of highly dispersed Fe(II)-doped g-C3N4 and its application in Fenton-like catalysis. Mol. Catal. 2017, 435, 156–165. [Google Scholar] [CrossRef]
  52. Guzmán, H.; Farkhondehfal, M.A.; Tolod, K.R.; Hernández, S.; Russo, N. Chapter 11-Photo/electrocatalytic hydrogen exploitation for CO2 reduction toward solar fuels production. In Solar Hydrogen Production; Calise, F., D’Accadia, M.D., Santarelli, M., Lanzini, A., Ferrero, D., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 365–418. [Google Scholar]
  53. Choi, W. Pure and modified TiO2 photocatalysts and their environmental applications. Catal. Surv. Asia 2006, 10, 16–28. [Google Scholar] [CrossRef]
  54. Li, X.; Yu, J.; Jaroniec, M.; Chen, X. Cocatalysts for Selective Photoreduction of CO2 into Solar Fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef]
  55. Viswanathan, B. Energy Sources: Fundamentals of Chemical Conversion Processes and Applications; Elsevier Science: Amsterdam, Netherlands, 2016. [Google Scholar]
  56. Wang, Z.; Li, C.; Domen, K. Recent developments in heterogeneous photocatalysts for solar-driven overall water splitting. Chem. Soc. Rev. 2019, 48, 2109–2125. [Google Scholar] [CrossRef] [PubMed]
  57. Di, L.; Yang, H.; Xian, T.; Chen, X. Enhanced Photocatalytic Degradation Activity of BiFeO3 Microspheres by Decoration with g-C3N4 Nanoparticles. Mater. Res. 2018, 21, 1–10. [Google Scholar] [CrossRef] [Green Version]
  58. Matos, J.; Arcibar-Orozco, J.; Poon, P.S.; Pecchi, G.; Rangel-Mendez, J.R. Influence of phosphorous upon the formation of DMPO-OH and POBN-O2•− spin-trapping adducts in carbon-supported P-promoted Fe-based photocatalysts. J. Photochem. Photobiol. A 2020, 391, 112362. [Google Scholar] [CrossRef]
  59. Matos, J.; Poon, P.S.; Montaña, R.; Romero, R.; Gonçalves, G.R.; Schettino, M.A., Jr.; Passamani, E.C.; Freitas, J.C.C. Photocatalytic activity of P-Fe/activated carbon nanocomposites under artificial solar irradiation. Catal. Today 2020, 356, 226–240. [Google Scholar] [CrossRef]
  60. Matos, J.; Rosales, M.; Garcia, A.; Nieto-Delgado, C.; Rangel-Mendez, J.R. Hybrid photoactive materials from municipal sewage sludge for the photocatalytic degradation of methylene blue. Green Chem. 2011, 13, 3431–3439. [Google Scholar] [CrossRef]
  61. Casado, J.; Fornaguera, J. Pilot-Scale Degradation of Organic Contaminants in a Continuous-flow Reactor by the Helielectro-Fenton Method. Clean 2008, 36, 53–58. [Google Scholar] [CrossRef]
  62. Sayama, K.; Yoshida, R.; Kusama, H.; Okabe, K.; Abe, Y.; Arakawa, H. Photocatalytic decomposition of water into H2 and O2 by a two-step photoexcitation reaction using a WO3 suspension catalyst and an Fe3+/Fe2+ redox system. Chem. Phys. Lett. 1997, 277, 387–391. [Google Scholar] [CrossRef]
  63. Khaing, K.K.; Yin, D.; Xiao, S.; Deng, L.; Zhao, F.; Liu, B.; Chen, T.; Li, L.; Guo, X.; Liu, J.; et al. Efficient Solar Light Driven Degradation of Tetracycline by Fe-EDTA Modified g-C3N4 Nanosheets. J. Phys. Chem. C 2020, 124, 11831–11843. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the g-C3N4 photocatalysts.
Figure 1. XRD patterns of the g-C3N4 photocatalysts.
Photochem 01 00030 g001
Figure 2. FT-IR spectra of the g-C3N4 photocatalysts.
Figure 2. FT-IR spectra of the g-C3N4 photocatalysts.
Photochem 01 00030 g002
Figure 3. UV-Vis diffuse reflectance (inset) and plots of (F × E)1/2 vs. photon energy of the g-C3N4 photocatalysts.
Figure 3. UV-Vis diffuse reflectance (inset) and plots of (F × E)1/2 vs. photon energy of the g-C3N4 photocatalysts.
Photochem 01 00030 g003
Figure 4. PL spectra of the g-C3N4 photocatalysts.
Figure 4. PL spectra of the g-C3N4 photocatalysts.
Photochem 01 00030 g004
Figure 5. Photocurrent generation of g-C3N4 photocatalysts recorded under 1 V external potential.
Figure 5. Photocurrent generation of g-C3N4 photocatalysts recorded under 1 V external potential.
Photochem 01 00030 g005
Figure 6. Mössbauer spectra of the samples CN-Fe0.3 (a) and CN-Fe1.1 (b), where the Fe2+ species are in green, the Fe3+ species are in blue, and the Fe2O3 subspectra are in red.
Figure 6. Mössbauer spectra of the samples CN-Fe0.3 (a) and CN-Fe1.1 (b), where the Fe2+ species are in green, the Fe3+ species are in blue, and the Fe2O3 subspectra are in red.
Photochem 01 00030 g006
Figure 7. Time dependence of yields of (a) CH4, (b) CO, and (c) H2.
Figure 7. Time dependence of yields of (a) CH4, (b) CO, and (c) H2.
Photochem 01 00030 g007
Figure 8. VB edge potentials measured by XPS.
Figure 8. VB edge potentials measured by XPS.
Photochem 01 00030 g008
Figure 9. Band structures of the g-C3N4 photocatalysts.
Figure 9. Band structures of the g-C3N4 photocatalysts.
Photochem 01 00030 g009
Table 1. BET surface area, average pore size, and total pore volume of the g-C3N4 photocatalysts.
Table 1. BET surface area, average pore size, and total pore volume of the g-C3N4 photocatalysts.
SampleBET Surface Area (m2/g)Average Pore Size (nm)Total Pore Volume (cm3/g)
CN1613.70.11
CN-Fe0.113190.12
CN-Fe0.31118.70.11
CN-Fe1.1738.50.12
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Edelmannová, M.; Reli, M.; Kočí, K.; Papailias, I.; Todorova, N.; Giannakopoulou, T.; Dallas, P.; Devlin, E.; Ioannidis, N.; Trapalis, C. Photocatalytic Reduction of CO2 over Iron-Modified g-C3N4 Photocatalysts. Photochem 2021, 1, 462-476. https://doi.org/10.3390/photochem1030030

AMA Style

Edelmannová M, Reli M, Kočí K, Papailias I, Todorova N, Giannakopoulou T, Dallas P, Devlin E, Ioannidis N, Trapalis C. Photocatalytic Reduction of CO2 over Iron-Modified g-C3N4 Photocatalysts. Photochem. 2021; 1(3):462-476. https://doi.org/10.3390/photochem1030030

Chicago/Turabian Style

Edelmannová, Miroslava, Martin Reli, Kamila Kočí, Ilias Papailias, Nadia Todorova, Tatiana Giannakopoulou, Panagiotis Dallas, Eamonn Devlin, Nikolaos Ioannidis, and Christos Trapalis. 2021. "Photocatalytic Reduction of CO2 over Iron-Modified g-C3N4 Photocatalysts" Photochem 1, no. 3: 462-476. https://doi.org/10.3390/photochem1030030

Article Metrics

Back to TopTop