Next Article in Journal
Construction of a Novel Ternary GQDs/g-C3N4/ZIF-67 Photocatalyst for Enhanced Photocatalytic Carbon Dioxide Reduction
Previous Article in Journal
Glucose Isomerization to Fructose Catalyzed by MgZr Mixed Oxides in Aqueous Solution
Previous Article in Special Issue
Heterogeneous Photocatalytic Oxidation and Detoxification of Simulated Agricultural Wastewater Contaminated with Boscalid Fungicide Using g-C3N4 Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing the Visible Light Photocatalytic Activity of TiO2-Based Coatings by the Addition of Exfoliated g-C3N4

by
Ilias Papailias
1,*,
Nadia Todorova
1,
Tatiana Giannakopoulou
1,
Niki Plakantonaki
1,
Michail Vagenas
1,
Panagiotis Dallas
1,
George C. Anyfantis
2,
Ioannis Arabatzis
2 and
Christos Trapalis
1,*
1
Institute of Nanoscience and Nanotechnology, NCSR “Demokritos”, Patriarchou Gregoriou E & 27 Neapoleos Str., 15341 Agia Paraskevi, Greece
2
NanoPhos S.A., Science and Technology Park of Lavrio, 19500 Lavrio, Greece
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(5), 333; https://doi.org/10.3390/catal14050333
Submission received: 28 March 2024 / Revised: 1 May 2024 / Accepted: 13 May 2024 / Published: 20 May 2024
(This article belongs to the Special Issue Recent Advances in g-C3N4-Based Photocatalysts)

Abstract

:
In the last few years, increasing interest from researchers and companies has been shown in the development of photocatalytic coatings for air purification and self-cleaning applications. In order to maintain the photocatalyst’s concentration as low as possible, highly active materials and/or combinations of them are required. In this work, novel photocatalytic formulations containing g-C3N4/TiO2 composites were prepared and deposited in the form of coatings on a-block substrates. The obtained photocatalytic surfaces were tested for NOx and acetaldehyde removal from model air. It was found that the addition of only 0.5 wt% g-C3N4 towards TiO2 content results in over 50% increase in the photocatalytic activity under visible light irradiation in comparison to pure TiO2 coating, while the activity under UV light was not affected. The result was related to the creation of a g-C3N4/TiO2 heterojunction that improves the light absorption and the separation of photogenerated electron-hole pairs, as well as to the inhibition of TiO2 particles’ agglomeration due to the presence of g-C3N4 sheets.

1. Introduction

The effect of outdoor and indoor pollutants on human health and the environment has been a matter of increasing importance during the last decades and therefore stricter legislations regarding emissions have been continuously applied. Among the most common pollutants, nitrogen oxides (NOx) [1,2], volatile organic compounds (VOCs) [3,4], and microorganisms [5,6] have been confirmed as significantly dangerous. Continuous exposure to them can cause serious respiratory problems, while at high concentrations they are considered particularly toxic. Construction materials in buildings are directly and constantly exposed to these kinds of pollutants [7], especially in enclosed spaces (such as workplaces, tunnels, underground parking, etc.) where a great accumulation of pollutants can occur. Importantly, these spaces’ surfaces can serve as large workable sites where the pollutants can be decomposed on. Thus, the research community is focusing on developing techniques and materials to utilize these surfaces for efficient air purification. In this direction, semiconductor photocatalysis is increasingly attracting the attention of researchers and companies since it is regarded as a green and viable technology that is efficient under normal ambient conditions [8]. After a photocatalyst is irradiated with UV or visible light, the generated electron-hole (e–h+) pairs are separated and can react with adsorbed water and/or oxygen from the atmosphere to form reactive oxygen species (ROS). Then, the ROS may react with the organic or inorganic molecules present on the surface of the photocatalyst, thus oxidizing and/or reducing them into harmless compounds. However, the photocatalysts cannot be used in their common powder form in real-life scenarios, as without proper immobilization they would drift away from the building surface. Therefore, photocatalytic building materials are fabricated by embedding the photocatalyst in the bulk of cements, paints, etc., or by applying the photocatalyst in the form of coatings on various surfaces [9] with the latter gaining increasing attention.
Considering that the photocatalytic reactions occur on the material surface and require direct contact between the photocatalyst and the pollutants, methods such as spray coating [10,11] and dip coating [12] need to be used to uniformly apply the photocatalytic material onto the substrate surface. In addition, to maintain the transmittance or the original color of the substrate, the coating layer should be highly transparent [13]. This means that the coating has to be thin and the concentration of the photocatalyst as low as possible. Furthermore, the coating deposition on existing building surfaces should not be followed by any severe post-deposition treatment such as high-temperature calcination, irradiation, etc. Ideally, the coating should be deposited by a one-step procedure at normal conditions (temperature and pressure) to have long-lasting bonding with the building surface and facile maintenance [14].
Titanium dioxide (TiO2)-based coatings have been studied and used for air purification and self-cleaning applications [14,15,16,17]. Undoubtedly, TiO2 is currently the most used photocatalyst in photocatalytic construction materials due to its high chemical stability and relatively low price [18,19]. Nonetheless, TiO2 effectiveness is limited by its UV responsiveness owing to its wide band gap (Eg = 3.2 eV) and fast e–h+ recombination caused by the high density of trap states [20,21]. Specifically, for the air depollution from NOx (NO and NO2) gases, novel photocatalysts are searched for in order to tackle the main disadvantages of the TiO2, i.e., UV sensitization and substantial release of toxic NO2 product. Other metal oxides [22,23], inorganic substances like Ag3PO4 [24], and organic polymers like g-C3N4 [25,26] alone or in the form of composites have been reported as very efficient photocatalytic materials owing to their visible light sensitization and/or low emission of NO2.
Amongst the newly developed materials, graphitic carbon nitride (g-C3N4) has emerged as a non-toxic, visible-light-driven (Eg = 2.7 eV) photocatalyst with exceptional optical properties and stability [27]. Over the past decade, g-C3N4 has been studied for renewable energy utilization [28] and environmental depollution. g-C3N4 has been proven promising for use in construction materials and coatings [29,30]; however, the expected wide practical application of the g-C3N4 photocatalyst has not been reached yet, neither as powder nor in the form of a coating.
While g-C3N4/TiO2 photocatalytic powders have been extensively investigated, g-C3N4/TiO2 in the form of a coating is still poorly explored. Some reports on solar fuel photocatalytic production [31], water purification [32], and air treatment [33] can be mentioned here. Generally, the increase in the photocatalyst’s amount improves the photocatalytic performance of coatings, however, their transmittance and homogeneity decrease [34]; thus, the great challenge remains on how to resolve this problem without compromising the optical and photocatalytic properties.
In this work, novel photocatalytic coatings containing g-C3N4/TiO2 composites were prepared by adding small quantities of chemically exfoliated g-C3N4 to a commercially available TiO2 suspension of NanoPhos S.A. [35]. The modified suspensions were then sprayed on a-block substrates and the obtained coatings were examined for outdoor and indoor air purification, specifically for photocatalytic removal of NOx and acetaldehyde pollutants under UV and visible light irradiation. The photocatalytic stability of the coatings was tested in order to evaluate their suitability for real-life applications; also, the physicochemical properties of the photocatalytic component, i.e., g-C3N4/TiO2 composites in the form of powder, were examined.

2. Results and Discussion

2.1. Crystalline Structure and Chemical Composition

The XRD patterns of the obtained powder samples are presented in Figure 1. The pattern of chemically exfoliated g-C3N4 is also given for comparison. All powder samples showed the characteristic for TiO2 peaks attributed to its anatase (ICDD, PDF-2 #21-1272) and rutile (ICDD, PDF-2 #21-1276) phases [36]. The patterns of the composite materials showed no distinct diffraction peaks related to the exfoliated g-C3N4. This can be ascribed to (i) its low content, i.e., 0.1, 0.5, 1.0, 2.0, and 5 wt% towards the TiO2; (ii) its low crystallinity due to the performed chemical exfoliation; and (iii) to the fact that the main (002) diffraction peak of g-C3N4 at 27.8°—corresponding to the interlayer stacking of aromatic rings with a distance of d = 0.32 nm [27]—overlaps with the (110) diffraction peak of the rutile phase at 27.4° [37].
The presence of the g-C3N4 in the composites was revealed by the FT-IR spectra shown in Figure 2. Specifically, the peaks between 1230 and 1630 cm−1 are associated with the characteristic stretching of C-N heterocycles including trigonal N-(C)3 and bridging H-N-(C)2 units [38,39]. Notably, the peaks become more prominent with the increase in the g-C3N4 concentration in the composites. Also, the broad peak from 400 to 700 cm−1 is attributed to the Ti-O stretching and Ti-O-Ti bridging stretching modes of TiO2 [40], thus confirming the presence of both components of the g-C3N4/TiO2 heterostructures. The observed broad peaks in the region from 3000 to 3500 cm−1 are attributed to N-H and O-H bonds originating from residual amino groups of the melamine precursor [41] and surface adsorbed water [42], respectively.

2.2. Surface Area and Morphology

The liquid N2 adsorption–desorption isotherms and the pore size distribution of the powder samples are shown in Figure 3a,b, correspondingly. All materials demonstrated characteristic hysteresis loops of type IV isotherms typical for mesoporous materials [43]. Regarding the pure components, the sample p-TiO2 showed a specific surface area (SSA) of 48.2 m2/g, while the exfoliated g-C3N4 exhibited a significantly higher SSA of 99.3 m2/g. The g-C3N4/TiO2 heterostructures exhibited SSA close to this of p-TiO2, which slightly increases with the increase in the g-C3N4 content. Namely, samples p-TiCN0.1, p-TiCN0.5, p-TiCN1, p-TiCN2, and p-TiCN5 showed SSA of 48.4, 48.6, 48.9, 49.1, and 49.2 m2/g, respectively. The materials demonstrated similar pore size distribution with a broad peak at ~180 Å (radius 18 nm) typical for the commercial P25 TiO2 [44,45]. The peak is indicative of small and large mesopores that correspond to pores within individual nanoparticles and spaces between the formed agglomerates, respectively [46]. The exfoliated g-C3N4 exhibited wider pore size distribution, which can be related to its layered structure.
In Figure 4, two typical SEM images of the samples p-TiO2 and p-TiCN5 are presented. For sample p-TiO2, the characteristic morphology of TiO2 particles forming large agglomerates can be observed. In sample p-TiCN5, the lamellar sheets of exfoliated g-C3N4 can also be seen intermingled with the TiO2 agglomerates. In this case, the TiO2 agglomerates appear to be smaller in size. This is a key finding since the g-C3N4 sheets seem to prevent the excessive agglomeration of TiO2 particles. This is expected to play an important role in the overall photocatalytic activity by improving the homogeneity of the prepared formulation suspension and the deposited on the a-block photocatalytic coating.

2.3. Light Absorbance and Band Gap Estimation

The measured diffuse reflectance spectra (R) of the powder samples, as well as the spectrum of g-C3N4, are shown in Figure 5a. It is evident that the g-C3N4 exhibited absorbance in the visible region (between 390 and 420 nm), while the p-TiO2 and all the composite g-C3N4/TiO2 samples have similar spectra, with the main absorbance positioned between 360 and 400 nm. By utilizing the Kubelka–Munk functions F = (1 − R)2/2R, the corresponding absorption functions (F × E)n = f(E) were constructed, as depicted in Figure 5b. The two semiconductors are considered to possess an indirect band gap, and to that end, n = 1/2 was used. The band gap energy (Eg) of the materials was estimated by extrapolation of the linear part of the absorption functions to the x-axis that is the photon energy in eV [47]. It can be seen that the Eg of the g-C3N4 is 2.98 eV due to its exfoliated nature, while the Eg of the p-TiO2 (3.05 eV) and all the composite g-C3N4/TiO2 samples are similar (3.02–3.03 eV). The results prove that the addition of g-C3N4 in low concentrations does not significantly affect the Eg of the materials.

2.4. Photocatalytic Activity

The variation in the gases’ (NO, NO2, and NOx) concentrations over the deposited photocatalytic coatings under dark and during light irradiation can be seen in the recorded experimental concentration curves in Figure S1. In Figure 6, the calculated changes in the gas concentrations, i.e., the decrease in the NO concentration, the increase in the NO2 concentration, and the total decrease in the NOx concentration in the air with regard to their initial concentrations (CNO = 1 ppm, CNO2 = 0 ppm, and CNOx = 1 ppm) are presented. Under UV light irradiation (Figure 6a), all coatings showed comparable photocatalytic activity with a slightly better performance of the coating with 0.5 wt% g-C3N4 (sample c-TiCN0.5) achieving 48% NOx removal. However, under visible light irradiation (Figure 6b), the addition of g-C3N4 led to significantly enhanced activity of the coatings. The coating with 1 wt% g-C3N4 (sample c-TiCN1) reached 12% total NOx removal, which is 4.5% higher than the coating with pure TiO2 (sample c-TiO2). This increase designates the important contribution of g-C3N4 in broadening the light response from UV to the visible light region [37], which is essential for the real-life application of the photocatalyst.
The results for the photocatalytic acetaldehyde oxidation by the coatings under UV and visible light irradiation are presented in Figure 7, while the recorded experimental concentration curves can be seen in Figures S2 and S3. It can be observed that the sample c-TiCN0.5 exhibited the best photocatalytic activity under both UV and visible light irradiation, reaching 57% and 7.5% acetaldehyde removal, respectively. Again, the difference was more prominent under visible light where the addition of 0.5 wt% g-C3N4 (sample c-TiCN0.5) led to ~4% higher removal than the coating with pure c-TiO2 (sample c-TiO2).
In the present study, it is shown that the addition of small quantities of g-C3N4 to TiO2-based formulations improves significantly the activity under visible light irradiation. Notably, major physicochemical properties influencing the photocatalytic efficiency, i.e., SSA and porosity, as well as light absorbance and width of the band gap were only slightly affected by the addition of g-C3N4. We would like to emphasize that the increased photocatalytic activity can be attributed to two factors: (i) the creation of a g-C3N4/TiO2 heterojunction that enhances the charge (e and h+) separation; and (ii) the reduced agglomeration of the TiO2 nanoparticles.
More specifically, the coupling between g-C3N4 and TiO2 (P25) investigated in our previous work [48] revealed that the positions of the band gap’s edges of the two semiconductors in the energy scale favor the formation of a heterojunction with band alignment type II. Then, taking into account the potentials for ·OH and ·O2 radicals’ formation, the most likely charge transfer mechanism is the Z-scheme [49,50], which facilitates the charge separation and reduces the recombination rate of photogenerated e–h+ pairs [51,52,53]. It should be noted here that, as it comprises two TiO2 phases (anatase and rutile), the P25 photocatalyst possesses its own charge separation mechanism [54,55], which results in enhanced formation of ·OH radicals on the nanoparticles’ surface. Through the Z-scheme, the g-C3N4 component provides an additional pathway for the e generated in the conduction band of the TiO2 component (CBTiO2) towards its valence band (VBg-C3N4) and finally to its conduction band (CBg-C3N4). Thus, the charge separation in the g-C3N4/TiO2 heterojunction and the formation of ·OH radicals on the composites’ surface are further enhanced. In relation to this, our previous EPR results on the formation of reactive oxygen species in g-C3N4/TiO2 systems with various ratios [56] also demonstrated that ·OH radicals are predominately formed in TiO2-rich systems under visible light irradiation. Considering also that NOx and acetaldehyde oxidation mechanisms are highly dependent on ·OH radicals, the enhanced photocatalytic removal of NOx and acetaldehyde recorded in this study can be attributed to the favorable charge transfer pathways and separation in the g-C3N4/TO2 heterostructures with low g-C3N4 content in comparison with pure p-TiO2.
The reduced agglomeration of the TiO2 nanoparticles—and therefore the improved homogeneity of the photocatalytic material in the formulation suspension—leads to the uniform distribution and better exposure of photocatalyst on the coated surfaces, facilitating the interaction of the gas pollutants with more photocatalytic centers. However, when the content of g-C3N4 exceeds a certain value (over 1 wt% towards TiO2), the activity of the photocatalytic coatings decreases. This occurs because TiO2 becomes covered by the g-C3N4 sheets, thus reducing the irradiation amount reaching the semiconductor. Also, according to [57,58], the photogenerated electron-hole pairs cannot reach the surface of the g-C3N4/TiO2 composite as the g-C3N4 layers become recombination centers. g-C3N4/TiO2 with 0.5–1.0 wt% g-C3N4 towards TiO2 appeared as the optimum ratios for a suitable balance of these processes, leading to higher photocatalytic activity of the composites in comparison to the pure TiO2 coating.
The finding that addition of only 0.5 wt% g-C3N4 towards TiO2 to a well-established commercial formulation of NanoPhos S.A. results in over 50% increase in the photocatalytic activity under visible light irradiation makes the coatings suitable for immediate practical application, especially for uses where the UV light is negligible (i.e., indoor applications), while at the same time, the production cost is kept as low as possible.
Being considered for real-life application, the c-TiCN0.5 sample was subjected to weathering in order to examine its long-term stability. The results shown in Figure 8 and Figure 9 demonstrate that after each 500-h cycle, the coating exhibited a decrease in photocatalytic activity, both for NOx and acetaldehyde removal, which is attributed to the wear and damage of the thin photocatalytic coating due to the intense weathering conditions. Critically though, even after three cycles of weathering, the coating remains active. The photocatalytic and weathering stability suggest a good adhesion of the coating to the a-block surface, which is extremely important for commercial application as no additional post-deposition treatment (usually necessary for the stabilization of coatings on a surface) is required. Considering the real-life application potential, it is established that the coating product maintains its value for approximately 18 to 36 months, which is well within the range of other commercial products.

3. Materials and Methods

3.1. Preparation of Materials and Coatings

Melamine powder was purchased from Alfa Aesar. Sulfuric acid (H2SO4, 98%) was purchased from Chem-Lab and isopropanol (IPA) from Sigma-Aldrich (Saint Louis, MO, USA). The chemicals were of analytical grade and were used without any further purification. The syntheses of bulk and chemically exfoliated g-C3N4 are described in detail in our previous work [59]. In brief, 4 g of melamine was put into an open alumina crucible and treated at 550 °C for 3 h in a tube furnace under Argon flow. The resulting yellow bulk g-C3N4 was ground into powder and subjected to chemical exfoliation. Namely, 0.5 g of bulk g-C3N4 was stirred in 8 mL concentrated H2SO4 for 3 h. Then, 2 mL distilled H2O was slowly added and the suspension was poured into 100 mL distilled H2O. After centrifugation and washing until pH 7, the remaining white powder was re-dispersed into 100 mL IPA and ultrasonicated for 1 h. Finally, the suspension was dried at 60 °C, thus obtaining exfoliated g-C3N4 powder.
The photocatalytic dispersions were prepared by adding a specific amount of exfoliated g-C3N4 in a commercially available TiO2-containing (2 wt%) formulation provided from NanoPhos S.A. The product is patented, while coatings deposited on a glass surface were examined in our previous work [16]. In this study, the amounts of the g-C3N4 were calculated, taking into account the known amount of TiO2 in the formulation suspension (Table 1). Quantities corresponding to 0.1, 0.5, 1, 2, and 5 wt% of g-C3N4 towards TiO2 were added and the modified suspensions were sonicated for 15 min using a 1000 W probe-type Hielscher UIP1000hdT ultrasonic device operating at 20 kHz (Wanaque, NJ, USA). Then, the suspensions were deposited on 5 cm × 10 cm × 0.5 cm a-block substrates (coated surface 50 cm2) using a Z-020-Tornador BLACK sprayer, (International Tool Company, Leicester, UK). The a-block is a lightweight, precast, foam-like concrete construction material with a porous surface. Its use is considered very beneficial as it insulates better than conventional bricks, while at the same time, it does not strain the building, reduces its load, and can withstand heavy constructions such as roofs. It should be mentioned that the coating was transparent and thin enough so that the coated and non-coated a-block substrates were similar in color (Figure 10). The coated substrates were left for curing at ambient conditions for at least 24 h before the photocatalytic tests. The photocatalytic coatings were named c-TiCNx, where c stands for coating and x is the wt% of g-C3N4 towards TiO2.
Due to the type of the substrate, physicochemical properties such as crystalline structure, specific surface area, light absorption, etc., were determined for the photocatalytic component alone. For this purpose, g-C3N4/TiO2 powders were obtained by drying 5 g of the suspensions in a ventilated oven at 105 °C for 48 h. Then, the solids were collected and ground into fine powders. The samples were named p-TiCNx, where p stands for powder and x is the wt% of g-C3N4 towards TiO2.

3.2. Characterization

A Siemens (Plano, TX, USA) D500 X-ray diffractometer was used for the XRD measurements. The FT-IR spectra of the materials were measured on a Nicolet (Wausau, WI, USA) iS50 spectrometer. The liquid N2 adsorption–desorption isotherms were obtained with a Quantachrome (Boynton Beach, FL, USA) Autosorb-iQ instrument. SEM characterization was performed using an FEI (Hillsboro, OR, USA) Quanta Inspect Microscope equipped with a tungsten filament operating at 25 kV. The UV–Vis diffuse reflectance spectra were recorded by a Shimadzu (Tokyo, Japan) UV-2100 spectrophotometer using BaSO4 as reference.

3.3. Photocatalytic Activity Evaluation

The photocatalytic activity of the coatings was measured for NOx removal from model gas and acetaldehyde oxidation under UV and visible light irradiation at room temperature. For the experiments, Philips (Andover, MA, USA) Cleo Compact 15 W lamps with an intensity of 10 W/m2 and Nordex (Chicago, IL, USA) T5-8W-4000 K lamps with an intensity of 7000 lux were used, respectively.

3.3.1. NOx Removal

The photocatalytic activity of the coatings for NOx removal was evaluated using a standard procedure based on ISO/DIS 22197-1 [60]. The setup configuration and experimental conditions are described in detail in our previous works [56,59]. Particularly, the prepared coated substrates were placed in a continuous flow reactor with dimensions 30 cm × 5 cm. The concentrations of NO, NO2, and NOx were monitored using a Horiba APNA-370 chemiluminescence-based NOx analyzer (with a lower detectable limit of 0.0005 ppm) equipped with an O3 lamp and a silicon photodiode sensor.

3.3.2. Acetaldehyde Oxidation

The photocatalytic activity of the coatings for acetaldehyde oxidation was evaluated using a standard procedure based on ISO/DIS 22197-2 [61]. Again, the setup configuration and experimental conditions are described in detail in our previous work [54]. The coated substrates were placed in a continuous flow reactor with dimensions 30 cm × 5 cm, while the concentration of acetaldehyde gas was monitored using a Shimadzu Tracera GC-2010 gas chromatograph equipped with a BID-2010 detector.

3.3.3. Photocatalytic Stability

The photocatalytic stability of the coatings was evaluated by performing consecutive weathering cycles using the QUV Accelerated Weathering Tester instrument by Q-Lab, which in a few days can reproduce the damage that occurs over months of outdoor exposure. To simulate outdoor weathering, the materials are subjected to alternating treatment with UV light (UVA-340 lamps at 0.71 W/m2), condensing humidity (at 50 °C), and water spray (with rate of 5 L/min), at controlled temperatures. It is estimated that a 500 h cycle corresponds to 6–12 months of real outdoor exposure depending on the geographic location and climate characteristics. In this work, 3 cycles of 500 h were performed and after each cycle the photocatalytic activity of the coating was measured.

4. Conclusions

Novel photocatalytic coatings were prepared by adding small quantities of exfoliated g-C3N4 in a commercially available TiO2 suspension and consequent deposition on construction material known as a-block. Under visible light irradiation, the coatings showed significantly increased photocatalytic activity in comparison to the commercial TiO2-containing coatings. The result was attributed to (i) the creation of a g-C3N4/TiO2 heterojunction that enhances the charge separation and reduces the recombination of the photogenerated electron-hole pairs; and (ii) the incorporation of g-C3N4, which inhibits the TiO2 particles’ agglomeration, thus facilitating the interaction of the gas pollutants with the photocatalytic centers. It was found that only 0.5–1.0 wt% of g-C3N4 towards the TiO2 content in the suspension is required, probably due to its exfoliated nature and high specific surface area. The fact that the color of the substrate is not affected and the production cost is kept low makes the coatings very attractive for practical implementation. Additionally, the coated a-block surfaces showed significant photocatalytic and weathering stability owed to the good adhesion of the coating to the substrate surface.
Overall, the developed photocatalytic coatings are very promising for real-life applications, as they display exceptional results with negligible extra cost and without the need for additional procedures. Hopefully, this work will inspire further research and contribute to the adoption of exfoliated g-C3N4 for commercial and practical applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14050333/s1. The following Figures are available: Figure S1. Experimental concentration curves of the monitored NO, NO2, and NOx gases under UV and visible light irradiation for the photocatalytic coatings; Figure S2. Experimental concentration curves of the monitored acetaldehyde gas under UV light irradiation for the photocatalytic coatings; Figure S3. Experimental concentration curves of the monitored acetaldehyde gas under visible light irradiation for the photocatalytic coatings.

Author Contributions

Conceptualization, I.P. and C.T.; Investigation, I.P., N.T., T.G., N.P., M.V., P.D. and G.C.A.; Resources, C.T. and I.A.; Supervision, C.T.; Writing—original draft and revision, I.P. and N.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the European Union and Greek national funds through the Attica Business Plan 2014–2020, action “Research and Innovation Cooperation in the Attica Region” (project code: ATTP4-0349671).

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

Authors George C. Anyfantis and Ioannis Arabatzis were employed by the company NanoPhos S.A., Science and Technology Park of Lavrio. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Lira, J.D.O.B.; Padoin, N.; Vilar, V.J.; Soares, C. Photocatalytic NOx abatement: Mathematical modeling, CFD validation and reactor analysis. J. Hazard. Mater. 2019, 372, 145–153. [Google Scholar] [CrossRef] [PubMed]
  2. Lasek, J.; Yu, Y.H.; Wu, J.C. Removal of NOx by photocatalytic processes. J. Photochem. Photobiol. C 2013, 14, 29–52. [Google Scholar] [CrossRef]
  3. Missia, D.A.; Demetriou, E.; Michael, N.; Tolis, E.I.; Bartzis, J.G. Indoor exposure from building materials: A field study. Atmos. Environ. 2010, 44, 4388–4395. [Google Scholar] [CrossRef]
  4. Kim, S.; Kim, J.A.; Kim, H.J.; Kim, S.D. Determination of formaldehyde and TVOC emission factor from woodbased composites by small chamber method. Polym. Test. 2006, 25, 605–614. [Google Scholar] [CrossRef]
  5. Veltri, S.; Palermo, A.M.; De Filpo, G.; Xu, F. Subsurface treatment of TiO2 nanoparticles for limestone: Prolonged surface photocatalytic biocidal activities. Build. Environ. 2019, 149, 655–661. [Google Scholar] [CrossRef]
  6. Goffredo, G.B.; Accoroni, S.; Totti, C.; Romagnoli, T.; Valentini, L.; Munafo, P. Titanium dioxide based nanotreatments to inhibit microalgal fouling on building stone surfaces. Build. Environ. 2017, 112, 209–222. [Google Scholar] [CrossRef]
  7. Wang, Z.; Gauvin, F.; Feng, P.; Brouwers, H.J.H.; Yu, Q. Self-cleaning and air purification performance of Portland cement paste with low dosages of nanodispersed TiO2 coatings. Constr. Build. Mater. 2020, 263, 120558. [Google Scholar] [CrossRef]
  8. Gartner, M.; Trapalis, C.; Todorova, N.; Giannakopoulou, T.; Dobrescu, G.; Anastasescu, M.; Osiceanu, P.; Ghita, A.; Enache, M.; Dumitru, L.; et al. Doped Sol-gel TiO2 Films for Biological Applications. Bull. Korean Chem. Soc. 2008, 29, 1038–1042. [Google Scholar]
  9. Yang, Y.; Ji, T.; Yang, Z.; Zhang, Y.; Su, W.; Wu, R.; Wu, Z. Efficiency and durability of g-C3N4-based coatings applied on mortar under peeling and washing trials. Constr. Build. Mater. 2020, 234, 117438. [Google Scholar] [CrossRef]
  10. Faraldos, M.; Kropp, R.; Anderson, M.; Sobolev, K. Photocatalytic hydrophobic concrete coatings to combat air pollution. Catal. Today 2016, 259, 228–236. [Google Scholar] [CrossRef]
  11. Lettieri, M.; Colangiuli, D.; Masieri, M.; Calia, A. Field performances of nanosized TiO2 coated limestone for a self-cleaning building surface in an urban environment. Build. Environ. 2019, 147, 506–516. [Google Scholar] [CrossRef]
  12. Li, Q.; Liu, Q.; Peng, B.; Chai, L.; Liu, H. Self-cleaning performance of TiO2-coating cement materials prepared based on solidification/stabilization of electrolytic manganese residue. Constr. Build. Mater. 2016, 106, 236–242. [Google Scholar] [CrossRef]
  13. Jiang, W.; Zong, X.; Wang, X.; Sun, Z. Transparent Coating with TiO2 Nanorods for High-performance Photocatalytic Self-cleaning and Environmental Remediation. Chem. Res. Chin. Univ. 2020, 36, 1097–1101. [Google Scholar] [CrossRef]
  14. Huang, Y.; Zhang, J.; Wang, Z.; Liu, Y.; Wang, P.; Cao, J.; Ho, W. g-C3N4/TiO2 composite film in the fabrication of a photocatalytic air-purifying pavement. Solar RRL 2020, 4, 2000170. [Google Scholar] [CrossRef]
  15. Todorova, N.; Giannakopoulou, T.; Pomoni, K.; Yu, J.G.; Vaimakis, T.; Trapalis, C. Photocatalytic NOx oxidation over modified ZnO/TiO2 thin films. Catal. Today 2015, 252, 41–46. [Google Scholar] [CrossRef]
  16. Arabatzis, I.; Todorova, N.; Fasaki, I.; Tsesmeli, C.; Peppas, A.; Li, W.X.; Zhao, Z. Photocatalytic, self-cleaning, antireflective coating for photovoltaic panels: Characterization and monitoring in real conditions. Sol. Energy 2018, 159, 251–259. [Google Scholar] [CrossRef]
  17. Vacaroiu, C.; Enache, M.; Gartner, M.; Popescu, G.; Anastasescu, M.; Brezeanu, A.; Todorova, N.; Giannakopoulou, T.; Trapalis, C.; Dumitru, L. The effect of thermal treatment on antibacterial properties of nanostructured TiO2(N) films illuminated with visible light. World J. Microbiol. Biotechnol. 2009, 25, 27–31. [Google Scholar] [CrossRef]
  18. Zhao, J.; Yang, X. Photocatalytic oxidation for indoor air purification: A literature review. Build. Environ. 2003, 38, 645–654. [Google Scholar] [CrossRef]
  19. Gauvin, F.; Caprai, V.; Yu, Q.; Brouwers, H. Effect of the morphology and pore structure of porous building materials on photocatalytic oxidation of air pollutants. Appl. Catal. B 2018, 227, 123–131. [Google Scholar] [CrossRef]
  20. Osborn, D.; Hassan, M.; Asadi, S.; White, J.R. Durability quantification of TiO2 surface coating on concrete and asphalt pavements. J. Mater. Civ. Eng. 2014, 26, 331–337. [Google Scholar] [CrossRef]
  21. Guan, S.; Hao, L.; Yang, Y.; Yoshida, H.; Zhao, X.; Lu, Y. Significantly enhanced photocatalytic activity of TiO2/TiC coatings under visible light. J. Solid State Electrochem. 2020, 25, 603–609. [Google Scholar] [CrossRef]
  22. Danish, M.S.S.; Estrella, L.L.; Alemaida, I.M.A.; Lisin, A.; Moiseev, N.; Ahmadi, M.; Nazari, M.; Wali, M.; Zaheb, H.; Senjyu, T. Photocatalytic applications of metal oxides for sustainable environmental remediation. Metals 2021, 11, 80. [Google Scholar] [CrossRef]
  23. Khanal, V.; Balayeva, N.O.; Günnemann, C.; Mamiyev, Z.; Dillert, R.; Bahnemann, D.W.; Subramanian, V.R. Photocatalytic NOx removal using tantalum oxide nanoparticles: A benign pathway. Appl. Catal. B Environ. 2021, 291, 119974. [Google Scholar] [CrossRef]
  24. Zhang, C.; Feng, C.; Yuan, J.; Wang, Z.; Wang, Y.; Zhou, S.; Gu, P.; Li, Y. Extended construction strategies of Ag3PO4-based heterojunction photocatalysts for robust environmental applications. J. Environ. Chem. Eng. 2023, 11, 110705. [Google Scholar] [CrossRef]
  25. Gu, Z.; Jin, M.; Wang, X.; Zhi, R.; Hou, Z.; Yang, J.; Hao, H.; Zhang, S.; Wang, X.; Zhou, E.; et al. Recent advances in g-C3N4-based photocatalysts for NOx removal. Catalysts 2023, 13, 192. [Google Scholar] [CrossRef]
  26. Li, Y.; Sun, Y.; Ho, W.; Zhang, Y.; Huang, H.; Cai, Q.; Dong, F. Highly enhanced visible-light photocatalytic NOx purification and conversion pathway on self-structurally modified g-C3N4 nanosheets. Sci. Bull. 2018, 63, 609–620. [Google Scholar] [CrossRef] [PubMed]
  27. 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]
  28. 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]
  29. Yang, Y.; Ji, T.; Su, W.; Yang, B.; Zhang, Y.; Yang, Z. Photocatalytic NOx abatement and self-cleaning performance of cementitious composites with g-C3N4 nanosheets under visible light. Constr. Build. Mater. 2019, 225, 120–131. [Google Scholar] [CrossRef]
  30. Peng, F.; Ni, Y.; Zhou, Q.; Kou, J.; Lu, C.; Xu, Z. New g-C3N4 based photocatalytic cement with enhanced visible-light photocatalytic activity by constructing muscovite sheet/SnO2 structures. Constr. Build. Mater. 2018, 179, 315–325. [Google Scholar] [CrossRef]
  31. Wang, L.; Fei, X.; Zhang, L.; Yu, J.G.; Cheng, B.; Ma, Y. Solar Fuel generation over nature-inspired recyclable TiO2/g-C3N4 S-scheme hierarchical thin-film photocatalyst. J. Mater. Sci. Technol. 2022, 112, 1–10. [Google Scholar] [CrossRef]
  32. Wei, Z.; Liang, F.; Liu, Y.; Luo, W.; Wang, J.; Yao, W.; Zhu, Y. Photoelectrocatalytic degradation of phenol-containing wastewater by TiO2/g-C3N4 hybrid heterostructure thin film. Appl. Catal. B Environ. 2017, 201, 600–606. [Google Scholar] [CrossRef]
  33. Zheng, M.-W.; Wen, W.-Y.; Liu, S.-H. g-C3N4/TiO2 for gas-phase formaldehyde photodegradation under visible light in the humidity control coatings. J. Taiwan Inst. Chem. Eng. 2024, 154, 105129. [Google Scholar] [CrossRef]
  34. Xi, R.; Wang, Y.; Wang, X.; Lv, J.; Li, X.; Li, T.; Zhang, X.; Du, X. Ultrafine nano-TiO2 loaded on dendritic porous silica nanoparticles for robust transparent antifogging self-cleaning nanocoatings. Ceram. Int. 2020, 46, 23651–23661. [Google Scholar] [CrossRef]
  35. Available online: https://nanophos.com (accessed on 20 March 2024).
  36. Tong, Z.; Yang, D.; Xiao, T.; Tian, Y.; Jiang, Z. Biomimetic fabrication of g-C3N4/TiO2 nanosheets with enhanced photocatalytic activity toward organic pollutant degradation. Chem. Eng. J. 2015, 260, 117–125. [Google Scholar] [CrossRef]
  37. Papailias, I.; Todorova, N.; Giannakopoulou, T.; Yu, J.G.; Dimotikali, D.; Trapalis, C. Photocatalytic activity of modified g-C3N4/TiO2 nanocomposites for NOx removal. Catal. Today 2017, 280, 37–44. [Google Scholar] [CrossRef]
  38. Thomas, A.; Fischer, A.; Goettmann, F.; Antonietti, M.; Mueller, J.O.; Schloegl, R.; Carlsson, J.M. Graphitic carbon nitride materials: Variation of structure and morphology and their use as metal-free catalysts. J. Mater. Chem. 2008, 18, 4893–4908. [Google Scholar] [CrossRef]
  39. Dong, F.; Wu, L.W.; Sun, Y.J.; Fu, M.; Wu, Z.B.; Lee, S.C. Efficient synthesis of polymeric g-C3N4 layered materials as novel efficient visible light driven photocatalysts. J. Mater. Chem. 2011, 21, 15171–15174. [Google Scholar] [CrossRef]
  40. Yu, J.G.; Yu, H.G.; Cheng, B.; Zhao, X.J.; Yu, J.C.; Ho, W.K. The effect of calcination temperature on the surface microstructure and photocatalytic activity of TiO2 thin films prepared by liquid phase deposition. J. Phys. Chem. B 2003, 107, 13871–13879. [Google Scholar] [CrossRef]
  41. Zhao, Y.; Liu, Z.; Chu, W.; Song, L.; Zhang, Z.; Yu, D.; Tian, Y.; Xie, S.; Sun, L. Large-scale synthesis of nitrogen-rich carbon nitride microfibers by using graphitic carbon nitride as precursor. Adv. Mater. 2008, 20, 1777–1781. [Google Scholar] [CrossRef]
  42. Bojdys, M.J.; Muller, J.-O.; Antonietti, M.; Thomas, A. Ionothermal synthesis of crystalline, condensed, graphitic carbon nitride. Chem. Eur. J. 2008, 14, 8177–8182. [Google Scholar] [CrossRef] [PubMed]
  43. 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] [PubMed]
  44. Fan, K.; Zhang, W.; Peng, T.; Chen, J.; Yang, F. Application of TiO2 fusiform nanorods for dye-sensitized solar cells with significantly improved efficiency. J. Phys. Chem. C 2011, 115, 17213–17219. [Google Scholar] [CrossRef]
  45. Kim, Y.E.; Byun, M.Y.; Lee, K.-Y.; Lee, M.S. Effects of chlorinated Pd precursors and preparation methods on properties and activity of Pd/TiO2 catalysts. RSC Adv. 2020, 10, 41462–41470. [Google Scholar] [CrossRef] [PubMed]
  46. Tao, T.; Glushenkov, A.M.; Chen, Q.; Hu, H.; Zhou, D.; Zhang, H.; Boese, M.; Liu, S.; Amal, R.; Chen, Y. Porous TiO2 with a controllable bimodal pore size distribution from natural ilmenite. CrystEngComm 2011, 13, 1322–1327. [Google Scholar] [CrossRef]
  47. 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]
  48. Giannakopoulou, T.; Papailias, I.; Todorova, N.; Boukos, N.; Liu, Y.; Yu, J.; Trapalis, C. Tailoring the energy band gap and edges’ potentials of g-C3N4/TiO2 composite photocatalysts for NOx removal. Chem. Eng. J. 2017, 310, 571–580. [Google Scholar] [CrossRef]
  49. Afroz, K.; Moniruddin, M.; Bakranov, N.; Kudaibergenov, S.; Nuraje, N. A heterojunction strategy to improve the visible light sensitive water splitting performance of photocatalytic materials. J. Mater. Chem. A 2018, 6, 21696–21718. [Google Scholar] [CrossRef]
  50. Li, J.; Zhang, M.; Li, Q.; Yang, J. Enhanced visible light activity on direct contact Z-schemeg-C3N4-TiO2 photocatalyst. Appl. Surf. Sci. 2017, 391, 184–193. [Google Scholar] [CrossRef]
  51. Zhou, C.; Ye, N.F.; Yan, X.H.; Wang, J.J.; Pan, J.M.; Wang, D.F.; Wang, Q.; Zu, J.X.; Cheng, X.N. Construction of hybrid Z-scheme graphitic C3N4/reduced TiO2 microsphere with visible-light-driven photocatalytic activity. J. Materiomics 2018, 4, 238–246. [Google Scholar] [CrossRef]
  52. Ma, J.; Wang, C.; He, H. Enhanced photocatalytic oxidation of NO over g-C3N4-TiO2 under UV and visible light. Appl. Catal. B Environ. 2016, 184, 28–34. [Google Scholar] [CrossRef]
  53. Lu, Z.; Zeng, L.; Song, W.; Qin, Z.; Zeng, D.; Xie, C. In situ synthesis of C-TiO2/g-C3N4 heterojunction nanocomposite as highly visible light active photocatalyst originated from effective interfacial charge transfer. Appl. Catal. B Environ. 2017, 202, 489–499. [Google Scholar] [CrossRef]
  54. Ohno, T.; Sarukawa, K.; Tokieda, K.; Matsumura, M. Morphology of a TiO2 photocatalyst (Degussa, P25) consisting of anatase and rutile crystalline phase. J. Catal. 2001, 203, 82–86. [Google Scholar] [CrossRef]
  55. Sun, B.; Smirniotis, P. Interaction of anatase and rutile TiO2 particles in aqueous photooxidation. Catal. Today 2003, 88, 49–59. [Google Scholar] [CrossRef]
  56. Papailias, I.; Todorova, N.; Giannakopoulou, T.; Dvoranova, D.; Brezova, V.; Dimotikali, D.; Trapalis, C. Selective removal of organic and inorganic air pollutants by adjusting the g-C3N4/TiO2 ratio. Catal. Today 2021, 361, 37–42. [Google Scholar] [CrossRef]
  57. Yu, J.G.; Wang, S.; Low, J.; Xiao, W. Enhanced photocatalytic performance of direct Z-scheme g-C3N4-TiO2 photocatalysts for the decomposition of formaldehyde in air. Phys. Chem. Chem. Phys. 2013, 15, 16883–16890. [Google Scholar] [CrossRef] [PubMed]
  58. Fagan, R.; McCormack, D.E.; Hinder, S.J.; Pillai, S.C. Photocatalytic Properties of g-C3N4-TiO2 Heterojunctions under UV and Visible Light Conditions. Materials 2016, 9, 286. [Google Scholar] [CrossRef]
  59. 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 Environ. 2018, 239, 16–26. [Google Scholar] [CrossRef]
  60. ISO 22197-1; Fine Ceramics (Advanced Ceramics, Advanced Technical Ceramics)—Test Method for Air-Purification Performance of Semiconducting Photocatalytic Materials, Part 1: Removal of Nitric Oxide. ISO: Geneva, Switzerland, 2016.
  61. ISO 22197-2; Fine Ceramics (Advanced Ceramics, Advanced Technical Ceramics)—Test Method for Air-Purification Performance of Semiconducting Photocatalytic materials, Part 2: Removal of Acetaldehyde. ISO: Geneva, Switzerland, 2019.
Figure 1. XRD patterns of the composite g-C3N4/TiO2 powder samples and the pure TiO2 and exfoliated g-C3N4.
Figure 1. XRD patterns of the composite g-C3N4/TiO2 powder samples and the pure TiO2 and exfoliated g-C3N4.
Catalysts 14 00333 g001
Figure 2. FT-IR spectra of the composite g-C3N4/TiO2 powder samples and the pure TiO2 and exfoliated g-C3N4.
Figure 2. FT-IR spectra of the composite g-C3N4/TiO2 powder samples and the pure TiO2 and exfoliated g-C3N4.
Catalysts 14 00333 g002
Figure 3. Liquid N2 adsorption–desorption isotherms (a) and pore size distribution curves (b) of the powder samples.
Figure 3. Liquid N2 adsorption–desorption isotherms (a) and pore size distribution curves (b) of the powder samples.
Catalysts 14 00333 g003
Figure 4. SEM images of powder samples p-TiO2 (a) and p-TiCN5 (b).
Figure 4. SEM images of powder samples p-TiO2 (a) and p-TiCN5 (b).
Catalysts 14 00333 g004
Figure 5. UV–Vis diffuse reflectance spectra (a) and plots of (F × E)1/2 vs. photon energy (b) of the powder samples.
Figure 5. UV–Vis diffuse reflectance spectra (a) and plots of (F × E)1/2 vs. photon energy (b) of the powder samples.
Catalysts 14 00333 g005aCatalysts 14 00333 g005b
Figure 6. Photocatalytic activity of the coatings for NOx removal under UV (a) and visible light (b) irradiation.
Figure 6. Photocatalytic activity of the coatings for NOx removal under UV (a) and visible light (b) irradiation.
Catalysts 14 00333 g006
Figure 7. Photocatalytic activity of the coatings for acetaldehyde oxidation under UV (a) and visible light (b) irradiation.
Figure 7. Photocatalytic activity of the coatings for acetaldehyde oxidation under UV (a) and visible light (b) irradiation.
Catalysts 14 00333 g007
Figure 8. Photocatalytic stability of the c-TiCN0.5 coating for NOx removal under UV (a) and visible light (b) irradiation, after three consecutive weathering cycles.
Figure 8. Photocatalytic stability of the c-TiCN0.5 coating for NOx removal under UV (a) and visible light (b) irradiation, after three consecutive weathering cycles.
Catalysts 14 00333 g008
Figure 9. Photocatalytic stability of the c-TiCN0.5 coating for acetaldehyde oxidation under UV (a) and visible light (b) irradiation, after three consecutive weathering cycles.
Figure 9. Photocatalytic stability of the c-TiCN0.5 coating for acetaldehyde oxidation under UV (a) and visible light (b) irradiation, after three consecutive weathering cycles.
Catalysts 14 00333 g009
Figure 10. Coated (left) and non-coated (right) a-block substrates.
Figure 10. Coated (left) and non-coated (right) a-block substrates.
Catalysts 14 00333 g010
Table 1. Experimental parameters for the preparation of the coatings: wt% of g-C3N4 towards TiO2, wt% of g-C3N4 in the suspension, and density (d) of the suspension.
Table 1. Experimental parameters for the preparation of the coatings: wt% of g-C3N4 towards TiO2, wt% of g-C3N4 in the suspension, and density (d) of the suspension.
Sampleg-C3N4 towards TiO2 (wt%)g-C3N4 in the Suspension (wt%)d (g/m3)
c-TiO2001.41
c-TiCN0.10.10.0021.42
c-TiCN0.50.50.0101.42
c-TiCN11.00.0201.42
c-TiCN22.00.0401.42
c-TiCN55.00.1001.43
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

Papailias, I.; Todorova, N.; Giannakopoulou, T.; Plakantonaki, N.; Vagenas, M.; Dallas, P.; Anyfantis, G.C.; Arabatzis, I.; Trapalis, C. Enhancing the Visible Light Photocatalytic Activity of TiO2-Based Coatings by the Addition of Exfoliated g-C3N4. Catalysts 2024, 14, 333. https://doi.org/10.3390/catal14050333

AMA Style

Papailias I, Todorova N, Giannakopoulou T, Plakantonaki N, Vagenas M, Dallas P, Anyfantis GC, Arabatzis I, Trapalis C. Enhancing the Visible Light Photocatalytic Activity of TiO2-Based Coatings by the Addition of Exfoliated g-C3N4. Catalysts. 2024; 14(5):333. https://doi.org/10.3390/catal14050333

Chicago/Turabian Style

Papailias, Ilias, Nadia Todorova, Tatiana Giannakopoulou, Niki Plakantonaki, Michail Vagenas, Panagiotis Dallas, George C. Anyfantis, Ioannis Arabatzis, and Christos Trapalis. 2024. "Enhancing the Visible Light Photocatalytic Activity of TiO2-Based Coatings by the Addition of Exfoliated g-C3N4" Catalysts 14, no. 5: 333. https://doi.org/10.3390/catal14050333

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