Biomimetic photocatalysts for the conversion of aqueous- and gas-phase nitrogen species to molecular nitrogen via denitrification and ammonia oxidation

Cheolwoo Park ab, Hyelim Kwak a, Gun-hee Moon *c and Wooyul Kim *a
aDepartment of Chemical and Biological Engineering, Research Institute of Global Environment, Sookmyung Women's University, 99, Cheongpa-ro 47-gil, Yongsan-gu, Seoul 04310, Republic of Korea. E-mail: wkim@sookmyung.ac.kr
bDepartment of Energy Science, Sungkyunkwan University, 2066, Seobu-ro, Jangan-Gu, Suwon, Gyeonggi-do 16419, Republic of Korea
cExtreme Materials Research Center, Korea Institute of Science and Technology (KIST), 5. Hwarang-ro 14-gil, Seongbuk-gu, Seoul 02792, Republic of Korea. E-mail: catalysis@kist.re.kr

Received 30th March 2021 , Accepted 30th June 2021

First published on 30th June 2021


Abstract

Denitrification and anaerobic ammonium oxidation (anammox) are important biological processes of the nitrogen cycle that help to preserve the global ecosystem. However, indiscriminate development and global population growth result in the discharge of large amounts of nitrogen species (e.g., via the Haber–Bosch process), particularly nitrogen oxides and ammonia, which cannot be fully digested by microorganisms and therefore accumulate in soil and water. Photocatalysts can promote the conversion of nitrogen oxides and ammonia to molecular nitrogen under the action of photogenerated electrons and holes, thus mimicking denitrifying and anammox bacteria, respectively. Herein, we review the biomimetic photocatalysts and photoelectrochemical cells used to convert aqueous and airborne nitrogen species to molecular nitrogen and shed light on the charge transfer mechanism that should be selectively controlled to favor the formation of molecular nitrogen over that of nitrogen-containing intermediates and by-products. Last but not least, we discuss the outlooks and perspectives of solar-powered molecular nitrogen recovery and suggest guidelines for the design of high-performance denitrification/anammox bacteria-like photocatalysts.


image file: d1ta02644e-p1.tif

Cheolwoo Park

Cheolwoo Park received a BS in Biochemical Engineering from Gangneung-Wonju National University in 2015 and is now pursuing a PhD in Energy Science under the supervision of Prof. Tae Kyu Ahn at Sungkyunkwan University (SKKU). He joined Sookmyung Women's University (advisor: Prof. Wooyul Kim) as a visiting researcher in 2016 and focuses on elucidating the reaction mechanisms of surface-modified metal oxide photocatalysts for energy conversion and environmental applications.

image file: d1ta02644e-p2.tif

Hyelim Kwak

Hyelim Kwak received a BS in Chemical and Biological Engineering from Sookmyung Women's University (Seoul, Korea) in 2020 and is currently pursuing a master's degree under the supervision of Prof. Wooyul Kim at Sookmyung Women's University, focusing on photocatalysis for environmental and energy applications.

image file: d1ta02644e-p3.tif

Gun-hee Moon

Gun-hee Moon received a BS in Chemical Engineering from Inha University (Incheon, Korea) in 2008, an MS degree in Environmental Engineering from POSTECH in 2011, and a PhD degree in Chemical Engineering from POSTECH (Pohang, Korea) in 2015. After a postdoctoral stay at the Max-Planck-Institut für Kohlenforschung (Germany, 2016–2020), he moved to the Korea Institute of Science and Technology (KIST), where he is currently employed as a senior scientist working on the design of photocatalysts, electrocatalysts, photoelectrochemical cells, and carbon materials for energy conversion and environmental applications.

image file: d1ta02644e-p4.tif

Wooyul Kim

Wooyul Kim received a PhD in Environmental Engineering (advisor: Prof. Wonyong Choi) from POSTECH (Pohang, Korea) in 2012 and worked at the Lawrence Berkeley National Laboratory (USA, 2012–2016) as a postdoctoral fellow (advisor: Dr Heinz Frei). During his PhD period, he joined Osaka University (advisor: Prof. Tetsuro Majima) as a visiting researcher in 2009, 2010, and 2012. He joined the faculty of the Department of Chemical and Biological Engineering at Sookmyung Women's University (Seoul, Korea) as an assistant professor (2016) and was promoted to associate professor in 2021. His current research focuses on revealing the structural identity of key intermediates and their roles in the catalytic cycle (i.e., their kinetic relevancy) to overcome the kinetic barriers for photo (or electro) catalysis. At an early stage of his independent career within the fields of photo(electro) catalysis for energy and environmental applications, he was honored as a winner of the 2020 Energy & Environmental Science lectureship award (Royal Society of Chemistry).


1. Introduction

Unlike ammonia (NH3), nitrite (NO2), and nitrate (NO3), molecular nitrogen (N2), which constitutes ∼80% of the atmosphere by volume, is not chemically and biologically reactive and cannot therefore be utilized by plants for the synthesis of proteins to be supplied to animals and humans.1,2 The conversion of N2 into more reactive forms mainly occurs through microbially facilitated processes such as nitrogen fixation and nitrification. The fixed reactive forms are absorbed, used, released in the form of excrements and dead bodies, converted back to the biologically inert form (i.e., N2) through denitrification and anammox, and eventually returned to the atmosphere.1 This biological nitrogen cycle preserved the balance between nitrogen fixation and denitrification for thousands of years.2 In the 20th century, however, human activities began to strongly impact the modern nitrogen cycle (established some 2.5 billion years ago), which is likely to result in the establishment of a new steady state in several decades.

Over the past century, the development of industrial processes and the rapid increase in fossil fuel use to satisfy the growing global demand for food and energy have drastically disrupted the nitrogen cycle.3–9 In particular, the Haber–Bosch process offers a way to synthetically fix nitrogen in the form of ammonia for the mass production of synthetic fertilizers,3 thus enabling abundant food production along with rapid world population growth. In the last 50 years, the consumption of fertilizers and fossil fuels has increased more than six-fold4,5 and three-fold,6 respectively. The projected global population growth is expected to result in elevated fertilizer production and fossil fuel consumption. As both of these anthropogenic sources (i.e., the Haber–Bosch process and fossil fuel combustion) account for ∼45% of the annual fixed nitrogen production (Fig. 1a), the above increase will initiate a cascade of large-scale environmental impacts such as (i) the extensive eutrophication of terrestrial and aquatic systems, (ii) the increase in potent greenhouse gas (i.e., N2O) inventory, and (iii) global acidification. The low (typically <40%) utilization efficiency of the nitrogen contained in fertilizers results in the nitrification-based conversion of large amounts of fertilizer (∼90% of NH4+) to highly mobile NO3 ions, which can leach into aquatic systems such as rivers, lakes, and aquifers.2 Moreover, besides producing N2 as the main product, anaerobic denitrification also affords N2O and thus significantly affects atmospheric N2O levels.7 As N2O reacts with the stratospheric ozone and is a potent greenhouse gas with global warming potential ∼300 times that of CO2, denitrification contributes to climate change and stratospheric ozone depletion.7 In addition, the absorption of nitrogen compounds by agricultural soils results in their acidification and thus inhibits the activity of soil organisms and disturbs the ecosystem. Fertilizer nitrogen is easily converted into gaseous ammonia and therefore returns from the atmosphere to the watershed via precipitation as another reactive nitrogen form.8 Nitrogen oxides (NOx) produced by fossil fuel combustion not only react with ammonia to produce fine dust and ozone and thus contribute to poor air quality, but also cause acid rain and, hence, soil and ocean acidification. The nitrogen cycle, the carbon cycle, and climate are known to exhibit numerous strong mutual interactions.9 The dramatic increase in atmospheric CO2 levels (>30% above pre-industrial values) due to fossil fuel combustion and land use change is viewed as the primary cause of climate warming observed over the past century. The human activity-induced perturbations of global nitrogen and carbon cycles are in part related to each other, as exemplified by the possible interacting drivers of these cycles during the 21st century (Fig. 1b).


image file: d1ta02644e-f1.tif
Fig. 1 (a) Rates of nitrogen flux in the modern nitrogen cycle depend on the efficiency of transformations between reservoirs. Reprinted with permission from ref. 2. Copyright© 2010, American Association for the Advancement of Science. (b) The main anthropogenic drivers of nitrogen–carbon–climate interactions in the 21st century. Reprinted with permission from ref. 9. Copyright© 2008, Nature Publishing Group.

From the perspective of nitrogen cycle management, the main objectives requiring special consideration are (i) the substantial decrease in nitrogen use,9 (ii) the direct up-cycling of used nitrogen to microbial protein,10 and (iii) the development of artificial denitrification processes powered by renewable energy. Among the various methods of decreasing nitrogen use, one can mention systematic crop rotation,10 optimization of fertilizer introduction timing and amount,11 and the breeding/development of genetically engineered crops for increasing nitrogen use efficiency.12 In view of the low efficiency of nitrogen utilization (e.g., agricultural nitrogen utilization efficiency = 40%, feed conversion efficiency = 15%, manure utilization efficiency = 50%),10 the direct up-cycling of used nitrogen to microbial protein has been proposed as an alternative to the formation of plant and meat proteins. As a renewable energy-powered direct denitrification process, the photocatalytic reduction of reactive nitrogen compounds to N2 holds great promise since the advantages of photocatalysis compared with conventional catalysis, thermocatalysis, and electrocatalysis are that (i) it does not require energy-intensive processes (solar energy vs. heat or electricity), (ii) the operation is possible without the need for oxidants, reductants, or electrolytes, (iii) it is flexible for application in both aqueous and gas-phase reactions, and (iv) material cost is relatively cheap.13–17 Herein, we introduce and discuss the most recent findings and advances in photocatalytic denitrification and ammonia oxidation processes, the ultimate goal of which is the conversion of reactive nitrogen compounds (e.g., ionic or gas-phase nitrogen oxides, ammonia) into inert N2 on a scale comparable to that of anthropogenic nitrogen fixation. Most parts of this review deal with denitrification/anammox bacteria-like photocatalysts and the related mechanisms, which greatly affect activity and selectivity.

2. Photocatalysts and co-catalysts for denitrification

2.1. Reduction of ionic nitrogen oxides to N2

The nitrogen cycle imbalance due to human activities (e.g., combustion, intensive fertilizer use, and agriculture) has resulted in increased nitrate levels in groundwater and other waters. Ionic nitrogen oxides including nitrate, nitrite, and NO are the prevalent contaminants in groundwater, causing eutrophication, water waste, and potentially health-threatening consequences such as cancer, birth defects, and cyanosis. Therefore, the release of ionic nitrogen oxides is strictly regulated, and their removal to secure water resources is a great challenge. Traditionally, the removal of NO3 from wastewater is achieved using reverse osmosis, ion exchange, and biological/catalytic treatment.1,18,19

Among the various denitrification methods used to reduce ionic nitrogen oxides in aquatic environments, eco-friendly photocatalytic denitrification is the one most promising from the perspective of industrialization.20–22 This photocatalytic reduction affords inert N2 gas and mainly involves the reduction of NO3 to N2via NO2.20,23–27 To increase the overall efficiency and selectivity for N2 formation, one needs to fully understand the underlying mechanisms, including those of the undesired nitrification (conversion of NO2 to NO3)28,29 and the dissimilatory nitrate reduction to ammonium (DNRA).30,31 In particular, these undesired reactions need to be precisely controlled (i.e., inhibited) to maximize N2 formation selectivity. However, from the perspective of reactive nitrogen up-cycling, the highly selective production of ammonium (e.g., DNRA) could be useful.32,33 In this section, we critically investigate the efforts made to enhance the selectivity for N2 production in the photocatalytic denitrification of ionic nitrogen oxides and the efficiency of this process. In particular, we demonstrate the important roles of intrinsic photocatalyst properties, sacrificial agents, and specific reaction conditions.

2.1.1. Photocatalytic materials for ionic nitrogen oxide reduction to N2. The discovery of the photocatalytic reduction of NO3 to NH4+ in aqueous solutions (over Pt–TiO2) in 1987[thin space (1/6-em)]34 triggered the search for other denitrification photocatalysts. In nature, the corresponding reaction involves multiple-electron transfer and is primarily performed by bacteria such as Thiobacillus denitrificans (2NO3 + 10e + 12H+ → N2 + 6H2O), whereas photocatalytic denitrification predominantly occurs in a stepwise manner and involves the reduction of NO3 to NO2 and the subsequent reduction of NO2 to N2 or NH4+.20,35–38 Each step of the nitrate-to-N2 reduction has its own rate constant, which is largely determined by the interaction between the photocatalyst and the adsorbed reactants (e.g., NO3, NO2, NO, and N2O) and reflects the ease of reactant adsorption on or product desorption from the photocatalyst.39 Attempts to accelerate photocatalytic denitrification have resulted in the discovery and evaluation of numerous photocatalysts, among which TiO2 is the one most frequently and thoroughly studied because of its durability, non-toxicity, long diffusion length of charge carriers, and other advantages. In particular, various methods of photocatalyst surface modification have been developed to accelerate photoconversion or alter the reaction mechanism and thus control selectivity. According to surface modifier type, these methods are classified into those relying on metal deposition,40,41 inorganic adsorbates,42,43 polymer coatings, dye sensitization,44–47 impurity doping, charge transfer complexation,48,49etc. Photocatalyst modification not only increases photoconversion efficiency but also affects selectivity by influencing the mechanism and kinetics of photocatalytic denitrification. Recent studies on photocatalysis by surface-modified semiconductors are summarized below.
Photocatalysts based on pristine TiO2 and related (bi)metallic composites. Compared to TiO2 powder (P25), TiO2 nanotubes (TNTs) achieved a ∼50% higher NO3 conversion efficiency and a slightly elevated selectivity for N2 formation, which was ascribed to their high specific surface area and abundant active sites.21 Given that bare TiO2 exhibits poor selectivity for N2 formation, TiO2–metal composites (M/TiO2) have been used for NO3 reduction in aquatic environments. The deposited metal changes the reactant adsorption properties and the charge carrier dynamics under illumination by modifying surface properties such as the space-charge region and charge density. In particular, metals deposited on TiO2 promote its use as an electron sink (with the formation of a Schottky barrier potential) by decreasing the work function and, hence, increasing the electron affinity.20et al. reported that the deposition of Cu, Fe, and Ag on TiO2 increased its activity and selectivity for N2, suggesting that the conversion efficiency and selectivity strongly depend on factors such as reaction temperature, hole scavenger presence/type, metal deposition method, and TiO2 particle size. Among the various M/TiO2 composites, Ag/TiO2 showed the highest activity to meet the EU-stipulated levels for drinking water (Table 1).50 The surface plasmon resonance effect of Ag0 and Au0 on TiO2 extended the excitation range of this oxide semiconductor from UV to visible and promoted the separation of photogenerated charge carriers (Fig. 2a).51 Compared to other M/TiO2 composites, Ag/TiO2 exhibited high selectivity for the reduction of NO3 to N2,21,40,41 while Au/TiO2 accelerated the formation of NH4+.51 The non-noble metals/TiO2 (Cu, Ni, Fe, Bi, Zn, etc.) composites also exhibited enhanced catalytic performance.33,52–54 For instance, Fe/TiO2 and Zn/TiO2 exhibited better N2 yield and selectivity than bare TiO2.54–56 Regarding Cu/TiO2 and Ni/TiO2, the formation of NO2 and NH4+ was predominant, where the nitrate reduction was facilitated by the electrons accumulated on Cu or Ni but the interaction with intermediates inhibited the generation of N2.33,50,54 On the other hand, Cr/TiO2 and Co/TiO2 lowered the conversion of NO3, which was explained by the light shield effect at a specific wavelength and the fast charge recombination kinetics as well.33,54 However, conventional metal-modified photocatalysts usually suffer from metal leaching, aggregation, and gradual deactivation and need to be substantially optimized in terms of N2 formation selectivity.
Table 1 Materials used for the photocatalytic reduction of ionic nitrogen oxides
No. Photocatalyst Co-catalyst Light Initial conc. Catalyst loading (g L−1) Sacrificial reagent NO3 conversion (%) N2 selectivity (%) By-products Ref.
1 TiO2 Medium-pressure Hg lamp, 150 W 1 mM 2.5 Oxalic acid 15 NH4+ 82
2 TiO2 Medium-pressure Hg lamp, 400 W 10 mM 10 Oxalic acid 9.8 56.5 NO2, NH4+ 33
3 TiO2 Medium-pressure Hg lamp, 150 W 0.8 mM 0.45 Oxalic acid 90.1 55.4 NO2, NH4+ 68
4 TiO2 High-pressure Hg lamp, 100 W 0.8 mM 0.38 Formic acid 48.5 38.1 NO2, NH4+ 80
5 KI 25.5 18
6 TiO2 High-pressure Hg lamp, 300 W 1.6 mM 1 Formic acid 26.8 72.4 NO2, NH4+ 81
7 TiO2 High-pressure Hg lamp, 150 W 1.6 mM 1 Formic acid 35.8 87.7 NO2, NH4+ 21
8 TiO2 (TNTs) 53.3 89.5 NO2, NH4+
9 TiO2 Au 400 W lamp 1.6 mM 0.21 Oxalic acid 44 NH4+ 83
10 TiO2 Cu High-pressure Hg lamp, 110 W 100 ppm 0.38 Formic acid 100 63 NH4+ 50
11 TiO2 Ag High-pressure Hg lamp, 300 W 1.6 mM 1 Formic acid 99.6 88.4 NO2, NH4+ 81
12 TiO2 Ag2O 97.5 82.9 NO2, NH4+
13 TiO2 (TNTs) AgCl High-pressure Hg lamp, 150 W 1.6 mM 1 Formic acid 94.5 92.9 NO2, NH4+ 21
14 TiO2 Pt–Cu High-pressure Hg lamp, 250 W 60 mg L−1 1 Benzene 59 ∼89 NO2, NH4+ 64
15 TiO2 Pd–Cu High-pressure Hg lamp, 400 W 0.05 mM 100 Formic acid 56 98 NO2 65
16 TiO2 Pd–Cu Medium-pressure Hg lamp, 150 W 1.6 mM 0.52 Formic acid 0.08 M 84 83 NO2, NH4+ 66
17 LiNbO3 High-pressure Hg lamp, 100 W 0.8 mM 0.38 Formic acid 98.4 95.8 NO2, NH4+ 80
18 High-pressure Hg lamp, 100 W 0.8 mM 0.38 KI 96.2 93 NO2, NH4+
19 LiNbO3 UV lamp 10 mg L−1 Membrane Formic acid 81.82 98.04 NO2, NH4+ 22
20 LiNbO3 High-pressure Hg lamp, 100 W 0.8 mM 0.4 Formic acid 60.5 57.21 NO2, NH4+ 76
21 Fe 86.69 85.71 NO2, NH4+
22 CuInS2 0.75 wt% Pt–0.75 wt% Ru Hg lamp, 125 W 7.2 mg L−1 0.5 Sodium oxalate 100 80.2 NO2 77
23 Xe lamp, 300 W (400 nm cut-off) 100 56.1 NO2
24 FeTiO3 Medium-pressure Hg lamp, 150 W 0.8 mM 0.45 Oxalic acid 100 93 NO2 68
25 GdCrO3 1 wt% Pd High-pressure Hg lamp, 500 W 0.8 mM 0.5 Formic acid 0.4 mM 98.7 100 67
26 1 wt% Ag 85.1 83.2 NO2, NH4+
27 1 wt% Cu 81.9 78.8 NO2, NH4+
28 79.3 81.4 NO2, NH4+
29 CuFe0.7Cr0.3S2 0.75 wt% Pd Hg lamp, 500 W 1.6 mM 1 Sodium oxalate 100 59 NO2 78
3 wt% Au
30 KTaO3 1 wt% Ni High-pressure Hg lamp, 450 W 10 mM 2.5 97 44 H2, NO2, NH4+ 69



image file: d1ta02644e-f2.tif
Fig. 2 Photocatalytic nitrate reduction promoted by (a) Au/TiO2 and (b) Pd–Cu/TiO2. Reprinted with permission from ref. 51 and 66. Copyright© 2011 and 2014, Elsevier. (c) Comparison of nitrate reduction promoted by LiNbO3 and P25 TiO2. Reprinted with permission from ref. 80. Copyright© 2016, American Chemical Society.

In bimetallic composites, the metals act as promoters and selectors. The promoter metal (e.g., Cu, Sn, In)57–59 initiates the rate-limiting step of the NO3 to NO2 conversion, while the selector metal (e.g., Pd, Pt, Rh)60,61 further reduces NO2 to NH4+ and/or N2. Among the available metal combinations, Pd–Cu is widely accepted as the most active and selective one for electrocatalytic NO3 reduction, the mechanism of which has been revealed by conventional electrochemical analysis and density functional theory (DFT) calculations.45,46 The increased H2 amount resulting from the elevated Pd loading and H2 flow rate promoted the reduction of CuII to Cu0 and thus facilitated NO3 removal, while the high N[thin space (1/6-em)]:[thin space (1/6-em)]H ratio on the active Pd sites increased the selectivity for N2.62 Bimetallic electrocatalysts have been widely deposited on photocatalysts for photocatalytic applications.63 Precious metal (e.g., Pt, Pd)–Cu combinations are among those offering the highest activity and selectivity for catalytic NO3 reduction (Table 1). Notably, NO3 was mainly converted to ammonia (over Pt/TiO2) or NO2 (over Cu/TiO2), whereas Pt–Cu/TiO2 catalysts exhibited a considerable selectivity for N2 formation in photocatalytic NO3 reduction.64 The fact that N2 formation was observed for Pd/TiO2 and Pd–Cu/TiO2 systems but was negligible for the Cu–TiO2 system means that Pd is indispensable for the photocatalytic reduction of NO2 to N2.65,66 Likewise, in bimetallic composites, electrons transferred from TiO2 to promoter metal sites reduce NO3 to NO2, with the subsequent reduction of NO2 to N2 occurring at selector metal sites. The adsorption of protons on the selector metal surface significantly affects the overall selectivity for N2 (Fig. 2b). Hence, the metal deposited on TiO2 controls the reaction path and, hence, the conversion efficiency and selectivity for N2 or NH4+, which implies that the optimization of the promoter-to-selector metal ratio is crucial for realizing selective N2 formation. Finally, the presence of sacrificial electron donors and the occurrence of competitive reactions (e.g., H2 production) present additional challenges.


New photocatalyst types. Much effort has been directed at the optimization of perovskite-based photocatalysts, as the unique properties of perovskites (e.g., chemical and optical stability, tunable bandgap and crystal structure, and long charge carrier lifetime) allow one to readily alter the dynamics of photogenerated charge carriers and the overall photocatalysis mechanism. Pd/GdCrO3 exhibited faster nitrate reduction and higher selectivity for N2 due to the negative conduction band energy level and the co-catalyst effect of Pd,67 while FeTiO3 was characterized by negligible NH4+ formation (i.e., it exhibited a remarkably high selectivity for N2) without the need for complex and expensive catalysts.68 KTaO3 effectively promoted the photocatalytic reduction of NO3 to NO2, N2, and NH3 under UV light irradiation even in the absence of co-catalysts or reducing reagents such as organic compounds.69

Layered double hydroxides (LDHs) with hydrotalcite-like structures are some of the interesting materials due to their unique properties such as anions intercalated in 2D interlayer spaces, a bunch of surface hydroxyl groups, flexibility to change elements, and swelling nature, where divalent (e.g., Mg, Co, Ni, Cu, and Zn) and trivalent (e.g., Al, Cr, Ni, and Ga) metal cations are combined.70,71 In particular, a high specific surface area, excellent electrical conductivity, high mobility of charge carriers, and high chemical stability make it possible to apply them in various photocatalytic reactions.72–74 Therefore, it was reported that the MgAl-LDH used for NO3 reduction enhanced the selectivity to N2 without any sacrificial reagent, which was ascribed to both attraction of NO3 ions near the photocatalyst surface and restriction of charge carrier recombination.75

LiNbO3 is a nonlinear optical material with high potential for NO3 removal,22,60,76 offering spontaneous polarization screening by either free electrons and holes or ions/molecules adsorbed on the surface. The second harmonic generation effects of nonlinear optical materials facilitate the generation of electrons and inhibit the recombination of charge carriers to enhance the efficiency and stability of NO3 reduction. The superior (compared to that of bare TiO2) activity of LiNbO3 was attributed to the photocatalytic reduction of nitrate through direct heterogeneous interactions with electrons at the conduction band of this material, whereas in the conventional photocatalysis mechanism, nitrate is mainly reduced by CO2˙ generated from holes at the valence band (Fig. 2c). In an effort to develop a systematic and durable industrial-scale process, LiNbO3 was applied to a membrane platform,22 which offered the inherent benefits of high separation performance and antifouling properties compared to common ultrafiltration membranes. In addition, LiNbO3 has been successfully applied to membrane materials without significant photocatalytic activity inhibition (Table 1). Fe–LiNbO3 exhibited an enhanced selectivity for N2 formation as well as a high NO3 conversion efficiency,76 which was ascribed to the increase in the specific surface area and the number of Lewis-acidic sites upon doping.

As wide-bandgap semiconductors (e.g., TiO2, FeTiO3, GdCrO3, and KTaO3) are intrinsic UV-light-driven photocatalysts, a more effective strategy would be to develop narrow-bandgap photocatalysts and thus utilize the whole solar spectrum. As a result, various chalcogenide materials (e.g., CuInS2[thin space (1/6-em)]77 and CuFe0.7Cr0.3S2[thin space (1/6-em)]78) have been developed. In particular, CuInS2 has a narrow bandgap of 1.45 eV and an insufficient conduction band potential for H2 production, thus preventing the over-reduction of nitrate to ammonia.

2.1.2. Insights into the mechanism of ionic nitrogen oxide reduction. The efficiency and selectivity of artificial solar denitrification systems can be increased by suppressing undesired reactions, which mainly correspond to DNRA (i.e., ammonification) and the re-oxidation of NO2 to NO3. In turn, a deep understanding of the overall mechanism is required to precisely control competitive reactions and thus selectively convert ionic nitrogen species to N2 while maintaining sufficient catalytic rates for keeping up with the photon flux at maximum solar intensity. DNRA is widely known as the anaerobic microbial pathway of the natural nitrogen cycle. The important implication of DNRA in denitrification is the production of NH4+ from NO3, which is the major side reaction for the reduction of NO3 to N2. The NO2 ion produced by the reduction of NO3 (NO3+ 2e + 2H+ → NO2 + H2O) can undergo ammonification (NO2 + 6e + 8H+ → NH4+ + 2H2O) or denitrification (NO2 + 6e + 8H+ → N2 + 4H2O), both of which are six-electron reductions. The preference for a particular NO2 reduction pathway is strongly affected by the environment, particularly by the ratio of N species to H atoms (N[thin space (1/6-em)]:[thin space (1/6-em)]H ratio) at active sites. The injection of excess H2 increases the surface concentration of H+ at active sites because of the low-energy-barrier dissociative adsorption of H2 and thus increases the selectivity for NH4+ formation (Fig. 3a). Alternatively, the N[thin space (1/6-em)]:[thin space (1/6-em)]H ratio can be changed by controlling the steady-state concentration of NO2 to increase the probability of binding of two nitrogen species to form N2 (Fig. 3b). The fine-tuning of the N[thin space (1/6-em)]:[thin space (1/6-em)]H ratio at surface active sites is required to suppress DNRA and thus selectively convert NO2 to N2.
image file: d1ta02644e-f3.tif
Fig. 3 Dependence of product (N2 and NH3) selectivity during NO2 reduction over Pd/TiO2 on (a) H2 flow rate and (b) NO2 concentration. Reprinted with permission from ref. 60. Copyright© 2014, American Chemical Society.

In the case of efficient nitrate conversion, photocatalytic nitrite oxidation, which is hard to detect during NO3 reduction, should be considered for low-efficiency nitrate reduction. Even if nitrification and denitrification occur simultaneously, it is difficult to identify the main factors of nitrification because of the same initial reactant and product. The formation of NO3 indicates that the oxidation of NO2 by holes occurs even in the presence of a hole scavenger in aqueous photocatalyst suspensions (Fig. 4). The low NO3 conversion efficiency is due to the low rate constant of NO3 reduction and the high rate constant of NO2 oxidation.


image file: d1ta02644e-f4.tif
Fig. 4 (a) Schematic mechanism of Pd/TiO2 operation and the combination of two photocatalytic systems for the reduction of NO2 to N2 and NO3. (b) Time-dependent conversion of NO2 and the formation of N2, H2, and NO3 in suspensions of Pd–TiO2 in aqueous sodium oxalate (diamonds: NO2, circles: N2, squares: NO3, triangles: H2). Reprinted with permission from ref. 29. Copyright© 2012, Royal Society of Chemistry.

The minimal loss of photogenerated electron–hole pairs offers flexibility for maximizing photocatalytic efficiency by adding sacrificial hole or electron scavengers to restrict charge carrier recombination. Sacrificial electron donors (i.e., hole scavengers) were used for photocatalytic denitrification in aqueous media to enable the efficient reduction of ionic nitrogen species5 and were shown to affect reactant–photocatalyst interactions. Sacrificial reagents act not only as efficient hole scavengers but also as precursors of active radicals for ionic nitrogen oxide reduction. Sacrificial reagents were demonstrated to promote the efficient removal of holes and thus reduce charge carrier recombination while being oxidized34 to afford strongly reducing (−1.81 vs. SHE) carboxyl radicals (CO2˙), which also resulted in activity enhancement.79 The most common sacrificial reagents are organic compounds such as formic acid,21,80,81 oxalic acid,68,82,83 humic acid,84 and methanol.34 Among them, formic acid is the best hole scavenger for NO3 reduction, as its simple structure results in the exclusive formation of the strongly reducing CO2˙, while the release of protons promotes efficient N2-selective reduction. Oxalic acid is the second most used hole scavenger, featuring a higher selectivity for NH4+ formation than formic acid (Table 1). The dependence of selectivity on hole scavenger type is attributed to the reduction ability of the reactant and intermediates. Simple carboxyl compounds (formic acid, sodium formate, etc.) are oxidized to afford abundant CO2˙ radicals and therefore allow for more efficient conversion than other organic hole scavengers. On the other hand, oxalic acid remarkably enhances the conversion of nitrite to N2 while exhibiting a modest hole scavenging ability (Fig. 5a and b).21,85 The understanding of the complicated interactions between various intermediates and sacrificial reagents is particularly challenging in the case of photocatalytic denitrification. In addition, the presence of additional reagents such as SO42−, H2PO4, F, Cl, and HCO3 increases system complexity and therefore leads to hole blocking and, as a consequence, inhibits hole scavenger oxidation by promoting surface anionization (Fig. 5c).86


image file: d1ta02644e-f5.tif
Fig. 5 (a) Effects of organic hole scavengers on the reduction of NO3 over AgCl/TiO2 nanotubes. Reprinted with permission from ref. 21. Copyright© 2018, Royal Society of Chemistry. (b) Effects of organic hole scavengers on the reduction of NO2 over Ag/TiO2. Reprinted with permission from ref. 85. Copyright© 2007, American Chemical Society. (c) Photocatalytic nitrate conversion efficiency and N2 selectivity achieved over TiO2 in the presence of different levels of SO42−, HCO3, H2PO4, F and Cl (irradiation duration: 3 h, sacrificial agent: formic acid). Reprinted with permission from ref. 86. Copyright© 2020, Elsevier.

2.2. Denitrification of airborne NOx and N2O to N2

Airborne nitrogen oxides (NOx) including NO and NO2 originate from anthropogenic (combustion of fossil fuels at high temperatures in automobile engines and power plants) and biogenic sources, significantly affecting human health and the environment.87 N2O is produced by human activities and natural processes and is both a greenhouse gas and an ozone-depleting substance.88 Taken together, NOx and N2O significantly affect the environment via (i) ozone and smog generation due to the reaction of NOx with volatile organic compounds (VOCs) upon irradiation with light, (ii) the acidification of water vapor (acid rain), (iii) excessive algal growth due to the dissolution of NOx in water (eutrophication), (iv) climate change and ozone layer depletion, and (v) secondary fine dust formation through the combination of the above oxides with water vapor, O3, ammonia, etc. Selective catalytic reduction (SCR)-based denitrification is a long-standing industrial process that is used to control air quality and is performed using different types of reactors and catalysts, depending on the application. Regarding the conventional deNOx process, ammonia has been used as a reducing agent instead of hydrogen to improve process safety and reduce the operating temperature.89 The conversion of NOx to N2 is described by Eley–Rideal and Langmuir–Hinshelwood models, which rely on the decomposition of intermediates generated through the reaction of surface-adsorbed activated NH3 with free gaseous NO and NO adsorbed on basic catalyst sites, respectively. Alternatively, airborne NOx can be removed via complete oxidation to NO3. Although this approach (nitrification) has received widespread attention for the development of smart cement and asphalt used in the construction of the state-of-the-art urban infrastructure,90 the accumulation of NO3 on the surface of photocatalysts can result in their deactivation. As this review focuses on denitrification only, readers interested in solar-driven nitrification are advised to consult appropriate literature.

The photocatalytic reduction of NOx and N2O to N2 offers the following advantages: (i) the use of water instead of the explosive H2 and the toxic NH3, (ii) operation at standard temperature and pressure, (iii) net zero carbon emission for operation under natural sunlight, and (iv) the availability of cheap and environmentally benign materials.91–94 As electron–hole pairs photogenerated in semiconducting materials can be transformed to the strongly oxidizing reactive oxygen species (ROS), aerobic conditions favor nitrification, in which case O2 acts as a good electron acceptor and a precursor of mobile hydroxyl radicals image file: d1ta02644e-t1.tif.95 Under anaerobic conditions, the electrons can be transferred to NOx or N2O without interference from O2 but can still be intercepted by water vapor (2H2O + 2e → H2 + 2OH) (Fig. 6). Ideally, the residual holes in photocatalysts oxidize water vapor (2H2O + 4h+ → O2 + 4H+), otherwise, nitrification is driven by oxidation with holes or OH˙ (H2O + h+ → OH˙ + H+). Along with selective charge transfer, other problems such as low solar light absorption, poor catalytic activity, need for noble metal–based co-catalysts, and lack of long-term durability should be overcome for practical applications. Herein, we summarize the strategies (e.g., structure and morphology control, co-catalyst loading, heteroatom doping, and hybridization with different types of materials) used to address the weaknesses of the photocatalytic reduction of NOx and N2O to N2.


image file: d1ta02644e-f6.tif
Fig. 6 Comparison of photocatalytic nitrification (left panel) and denitrification (right panel) mechanisms. The bottom right panel presents NH3-mediated denitrification, the release of trapped oxygen atoms from NO (green box), and the oxidation of sacrificial gases (purple box; ED stands for carbon-containing electron donors such as CO, CH4, C3H8, CH3OH, and C2H5OH).
2.2.1. Photocatalysts for the reduction of airborne NOx to N2.
DeNOx without additional electron donors. Water vapor and surface-trapped oxygen atoms formed by electron transfer to NO are ideal electron donors for the photocatalytic denitrification of NOx; however, the kinetics of water oxidation is sluggish, and the control of selective electron transfer to NOx is difficult. Therefore, whereas most studies focus on photocatalyst structure/surface modification and the operation of deNOx in the presence of nitrogen- or carbon-containing electron donors (e.g., NH3, CO, CH4, C3H8, CH3OH, and C2H5OH), very few processes have been performed without these donors. Wu et al. optimized the selectivity of NO reduction to N2 by controlling oxygen vacancies in TiO2 nanoparticles.96 These vacancies were introduced by doping the TiO2 lattice with Fe3+, and the substitution of Ti4+ by Fe3+ contributed to the charge compensation between negatively charged dopants and positively charged oxygen vacancies.97 In air, the photocatalytic conversion of NO over pristine TiO2 was not selective for N2 (Fig. 7a and Table 2), whereas a significantly improved selectivity was observed for Fe-doped TiO2 (Fig. 7b). The highest yield and selectivity for N2 were obtained under anaerobic conditions, and the formation of NO2 over Fe-doped TiO2 was completely suppressed (Fig. 7c). Thus, only N2 and O2 were detected along with unreacted NO in the outlet (Fig. 7d). The proposed mechanism (Table 3) is believed to involve ROS (superoxide anion radicals and hydroxyl radicals) as nitrification agents. On the other hand, oxygen vacancies stabilized by Fe doping enhanced selective charge transfer to NO, which did not involve radical species, providing a clue to the design of highly effective denitrification photocatalysts.
image file: d1ta02644e-f7.tif
Fig. 7 Photocatalytic conversion of NO under UV light irradiation in air over (a) pristine TiO2 and (b) Fe-doped TiO2. (c) Photocatalytic conversion of NO to N2 over Fe-doped TiO2 and the schematic mechanism of this conversion. The rapid decrease of NO in (a–c) is due to the formation of NO3via a reaction with superoxide anion radicals produced from adsorbed oxygen. (d) NO conversion ([NO] = 100 ppm) to N2 and O2 over pristine TiO2 and Fe-doped TiO2 in He. Reprinted with permission from ref. 96. Copyright© 2012, American Chemical Society.
Table 2 Performances of various materials as photocatalysts for the denitrification of airborne NOx
No. Photocatalyst Target (Conc.) Light Temp. Flow rate & GHSV Carrier gas Supplements NO conversion (%) N2 selectivity (%) By-products Ref.
1 TiO2 NO (1000 ppb) UV r.t. 1 L min−1 Air None ∼50 NO2, NO3 96
2 Fe-doped TiO2 Air ∼6 ∼50 NO2, NO3
3 N2 ∼4.5 ∼100 NO2, NO3
4 g-C3N4 NO (600 ppb) Visible (420 nm ≤) r.t. 1 L min−1 Air None ∼38 NO2, undefined 98
5 NO (1500 ppb) Ar Almost negligible
6 g-C3N4 with carbon vacancies NO (600 ppb) Air ∼48 NO2, undefined
7 NO (1500 ppb) Ar ∼34 ∼66 NO2
8 Cu+-ZSM-5 NO (2 Torr) UV (280 nm <) r.t. None ∼2 (4 h) ∼100 99
9 Ti-MCM-41 NO (180 μmol Eg-cat−1) UV (240 nm <) r.t. None ∼1.1 (1 h) ∼75 N2O, undefined 103
10 JCR-TiO2 (anatase) NO (1000 ppm) UV r.t. GHSV 32[thin space (1/6-em)]000 h−1 2% O2, 98% Ar NH3 (1000 ppm) 41 100 108
11 JCR-TiO2 (rutile) 53 100
12 JCR-TiO2 (anatase 91.3% + rutile 8.7%) 63 100
13 V2O5/TiO2 (1 wt%) NO (1000 ppm) UV r.t. GHSV 50[thin space (1/6-em)]000 h−1 2% O2, 98% Ar NH3 (1000 ppm) 17.7 100 114
14 CrO6/TiO2 (1 wt%) 34.2 100
15 MnO/TiO2 (1 wt%) 12.1 100
16 Fe2O3/TiO2 (1 wt%) 29.6 100
17 CoO/TiO2 (1 wt%) 21.6 100
18 NiO/TiO2 (1 wt%) 27.0 100
19 CuO/TiO2 (1 wt%) 26.1 100
20 ZnO/TiO2 (1 wt%) 46.6 100
21 Y2O3/TiO2 (1 wt%) 47.0 100
22 ZrO2/TiO2 (1 wt%) 41.1 100
23 Nb2O5/TiO2 (1 wt%) 58.4 >99
24 MoO3/TiO2 (1 wt%) 60.2 >99
25 Ta2O3/TiO2 (1 wt%) 38.6 100
26 WO3/TiO2 (1 wt%) 63.6 >99
27 N3-dye TiO2 NO (1000 ppm) Visible (420 nm ≤) r.t. GHSV 100[thin space (1/6-em)]000 h−1 2% O2, 98% He NH3 (1000 ppm) >99 >99 121
28 LaFe0.4Mn0.6O3/attapulgite NO (1000 ppm) UV r.t. GHSV 50[thin space (1/6-em)]000 h−1 3% O2, 97% N2 NH3 (1000 ppm) ∼85 ∼100 122
29 La0.7Ce0.3FeO3/attapulgite NO (1000 ppm) UV r.t. 0.1 L min−1 3% O2, 97% N2 NH3 (1000 ppm) ∼80 123
30 LaFe0.5Ni0.5O3/palygorskite NO (1000 ppm) Visible (420 nm ≤) 200 °C GHSV 50[thin space (1/6-em)]000 h−1 3% O2, 97% N2 NH3 (1000 ppm) ∼92 ∼98 124
31 La0.5Pr0.5CoO3/palygorskite NO (1000 ppm) Visible (420 nm ≤) 200 °C GHSV 40[thin space (1/6-em)]000 h−1 3% O2, 97% N2 NH3 (1000 ppm) ∼95 ∼99 125
32 N-doped carbon quantum dot-modified PrFeO3/palygorskite NO (1000 ppm) Visible 150–200 °C GHSV 50[thin space (1/6-em)]000 h−1 3% O2, 97% N2 NH3 (1000 ppm) ∼93 100 127
33 Fe2O3/SmFeO3/palygorskite NO (1000 ppm) UV <200 °C GHSV 40[thin space (1/6-em)]000 h−1 3% O2, 97% N2 NH3 (1000 ppm) ∼95 100 128
34 LaCoO3/appapulgite/rGO NO (1000 ppm) UV 100–150 °C GHSV 50[thin space (1/6-em)]000 h−1 3% O2, 97% N2 NH3 (1000 ppm) ∼95 100 129
35 Ag/TiO2 (1 wt%) NO (909 ppm) UV r.t. 5.5 cm3 min−1 Ar CO (1818 ppm) ∼35 μmol h g−1cat−1 90 136
36 Ag/TiO2 (5 wt%) ∼10 μmol h g−1cat−1 100
37 TiO2 NO (3000 ppm) UV 150 °C 5% O2, 3% H2O in N2 Carbon black 97 99 N2O 150


Table 3 Proposed mechanism for the conversion of NO on pristine and Fe-doped TiO2.96
Byproduct (i.e., NO2 and NO3) formation over TiO2
[Reductive pathway] [Oxidative pathway]
Superoxide radical-mediated Hydroxyl radical-mediated
a Numbers 1–4 denote the reaction pathway numbered in the scheme of Fig. 7c. image file: d1ta02644e-t17.tif and image file: d1ta02644e-t18.tif denote charged and neutral oxygen vacancies, respectively.
Ti(O2)ads + e → Ti(O2)ads Ti–OH + h+ → Ti–OH˙
Ti(O2)ads + NO(g) → Ti(NO3)ads Ti–OH˙ + NO(g) → Ti–H + NO2(g)
Ti(O2)ads + Ti–OH + hv + 2NO(g) → Ti(NO3)ads + Ti–H + NO2(g)

Conversion of NO to N2 over Fe-doped TiO2
[Reductive pathway] [Oxidative pathway]
image file: d1ta02644e-t15.tif (1) image file: d1ta02644e-t16.tif (4)
2Osurf–N → 2Osurf + N2(g) (2)
2Osurf → O2(g) (3)
2NO(g) + 4hv → N2(g) + O2(g)


Likewise, according to Dong et al.,98 the carbon vacancy tailoring of graphitic carbon nitride (g-C3N4) nanosheets (Cv-g-C3N4) significantly enhanced NO reduction under visible light irradiation (Fig. 8a). This reduction was faster in air than in argon irrespective of structure modification, as the ROS-mediated oxidation of NO to NO2 is much more favorable under oxic conditions (Fig. 8b). While the photocatalytic removal of NO over g-C3N4 was almost prohibited under anaerobic conditions, Cv-g-C3N4 was characterized by a relatively high conversion of NO and high selectivity for N2 formation (Fig. 8c), as in the case of Fe-doped TiO2. In line with the enhanced light absorption of Cv-g-C3N4 and the restricted recombination of charge carriers therein, electron spin resonance (ESR) spectra suggested that surface adsorption sites stimulated the chemisorption of airborne NO (via the interaction between a carbon atom with an unpaired electron in g-C3N4 and a nitrogen atom with an unpaired electron in NO), with carbon vacancies acting as active centers to induce interactions with the NO oxygen (i.e., Cv–O–N). Thus, after the adsorption of NO on g-C3N4 and Cv-g-C3N4, the ESR signal intensity due to carbon atoms with unpaired electrons decreased and increased, respectively (Fig. 8d and e). Illumination of Cv-g-C3N4 with pre-adsorbed NO induced a peak shift and the appearance of two new peaks at 3535 and 3555 G, which indicated the change of defect sites and the decomposition of NO into atomic N and O, respectively (Fig. 8f).


image file: d1ta02644e-f8.tif
Fig. 8 (a) Photocatalytic removal of NO ([NO] = 1500 ppb) in air and argon over g-C3N4 and Cv-g-C3N4 under UV light irradiation. Production of (b) NO2 in air and (c) N2 in argon over g-C3N4 and Cv-g-C3N4. ESR spectra of (d) g-C3N4 and (e) Cv-g-C3N4 before and after NO adsorption. (f) ESR spectra of Cv-g-C3N4 with adsorbed NO recorded in the dark and under UV light irradiation. Reprinted with permission from ref. 98. Copyright© 2017, Elsevier.

Prior to the strategies described above, Anpo et al. reported the UV light-promoted decomposition of NO into N2 and O2 over cation-exchanged ZSM-5 (Cu+, Ag+, and Pb2+),99–101 vanadium silicate (VS)/ZSM-5,102 and Ti-MCM-41.103 Herein, we do not discuss the characteristics of such materials and the corresponding kinetic analysis in detail because of the multitude of related reviews104–106 but briefly overview the concept of transition metal ion-mediated electron transfer to NO. Cu+ and Ag+ immobilized in zeolites can be excited under illumination and transfer an electron to the π-antibonding orbital of NO while concomitantly accepting the electron of another NO molecule. Consequently, two contiguous N⋯O species adsorbed at metal ion sites are converted into N2 and O2. Moreover, the coordination and distribution of metal oxide species, e.g., four-fold tetrahedrally coordinated vanadium oxide species with a terminal oxovanadium group (V[double bond, length as m-dash]O) in VS/ASM-5 and Ti oxide species with tetrahedrally coordinated Ti4+ in Ti-MCM-41, strongly affected NO removal activity and selectivity. Although several studies demonstrated the selective conversion of NO to N2 in the absence of supplements, the related yields were quite low, and the formation of undesired products could not be avoided, which was ascribed to catalyst inactivation via product accumulation on active sites. Another way to overcome this issue is the utilization of NH3 and carbon-containing compounds as sacrificial molecules.


DeNOxvia NH3(g) oxidation. The photo-assisted SCR (photo-SCR) of NO(g) with NH3(g) is an effective way to return N2 to the atmosphere (4NO + 4NH3 + O2 → 4N2 + 6H2O). Back in 1992, Cant et al. reported photo-SCR based on the reaction of NO with NH3 over TiO2 under UV light irradiation and used isotope labeling to elucidate the reaction pathway (414NO + 415NH3 + O2 → 414N15N + 6H2O).107 In this system, NO removal was very slow, and N2O and NO3 were still formed; moreover, the experiment was conducted under unrealistic conditions. Over the past 20 years, Tanaka's group systematically investigated the photo-SCR of NO, mainly focusing on the synthesis of TiO2-based photocatalysts, reaction mechanism verification, and the design of flow-type photoreactors operable at high gas hourly space velocities (GHSVs).108,109 From a mechanistic point of view, the photo-SCR of NO over TiO2 photocatalysts is suggested to comprise five steps (Fig. 9a).110,111 According to this mechanism, NH3(g) is adsorbed on the Lewis-acidic sites of TiO2 (step 1) and is then oxidized by a photogenerated hole (step 2), while the photogenerated electron is trapped by Ti4+. The adsorbed image file: d1ta02644e-t2.tif reacts with NO(g) through the Eley–Rideal mechanism (step 3), and the nitrosoamide (NH2NO) formed as an intermediate decomposes into N2 and H2O (step 4). The remaining Ti3+ is oxidized by electron transfer to O2(g) to regenerate Ti4+ (step 5). As NH3 adsorption takes place on Lewis-acidic sites, surface acidity control is an effective way of improving photo-SCR performance, whereas non-acidic surface area and the crystal phase do not contribute to activity enhancement.112 The rate-limiting step is affected by the concentration of O2, corresponding to step 4 in the presence of excess O2 and step 5 at O2 contents of <2 vol%. Ji et al.113 suggested that (i) the adsorption of NH3 and its direct oxidation by photogenerated holes was preferred to the dissociative adsorption of NH3 on TiO2, as proton-coupled hole transfer was energetically favored (NH3 + h+ + O2fimage file: d1ta02644e-t3.tif + O2fH; O2f denotes a two-fold coordinated O atom in TiO2) and the formation of NH2NO proceeded via the Eley–Rideal mechanism, and (ii) the decomposition of NH2NO into N2 and O2 was initiated by the transfer of Hb to the O atom, which was followed by either the transfer of Ha to O–Hb or the transfer of O–Hb to the surface Ti atom (Fig. 9b). Step 4 was calculated to have the highest energy barrier among other steps (i.e., it was the rate-limiting step), in line with the experimental result of Tanaka's group.
image file: d1ta02644e-f9.tif
Fig. 9 (a) Mechanism of the photocatalytic reduction of NO by NH3 over TiO2. Reprinted with permission from ref. 111. Copyright© 2004, Elsevier. (b) Potential energy diagram for the decomposition of NH2NO on the (101) surface of anatase TiO2. Reprinted with permission from ref. 113. Copyright© 2014, American Chemical Society.

To realize high catalytic performance or to run the system under visible light, one should appropriately design or modify the photocatalysts. Herein, photocatalysts were classified as those based on TiO2 or other materials. When TiO2 was modified with transition metal (V, Cr, Mn, Fe, Co, Ni, Cu, Zn, Y, Zr, Nb, Ta, and W) oxides, increased NO conversion was observed only for the more acidic ZnO, Y2O3, Nb2O5, MoO3, and WO3 (Fig. 10a and Table 2).114 The low activity observed for other oxides was ascribed to their non-photocatalytic nature or the instability of active sites. The highest activity of WO3/TiO2 was attributed to the facile decomposition of NH2NO on the weakly Lewis-acidic sites of WO3.115 The doping of Si into TiO2 caused the formation of smaller crystals with a higher surface area and pore volume, and acidity was enhanced because of the increased concentration of surface hydroxyl groups.116 The morphology of TiO2 was tailored by Ti foil anodization, and high-aspect-ratio TiO2 nanotubes provided more sites for NH3 adsorption than spherical TiO2 (P25).117 Although TiO2 does not absorb visible light, the adsorption of NH3 on TiO2 could generate an extra energy level in the bandgap via in situ doping to induce direct electron transfer from the N 2p orbital to the Ti 3d orbital under irradiation with visible light (λ ≥ 400 nm).118 This concept resembles that of ligand-to-metal charge transfer.119 Dye sensitization is an effective way to inject electrons from dye molecules into the conduction band of TiO2 under visible light. Among the 15 dyes anchored on TiO2, the Ru(2,2′-bipyridyl-4,4′-dicarboxylic acid)2(NCS)2 complex (N3-dye) showed the highest performance (Fig. 10b).120,121 The remaining holes in the HOMOs of dye molecules activated NH3, and N2 was selectively formed by the reaction between NO2 and image file: d1ta02644e-t4.tif. Consequently, the complete conversion of NO and a 100% selectivity for N2 were achieved at a high GHSV of 100[thin space (1/6-em)]000 h−1 under 30 min irradiation with visible light (Table 2). One of the serious problems of dye-sensitized systems in aqueous media is the detachment of dye molecules from TiO2 and the dependence of charge transfer on the complexation between functional groups. However, the occurrence of the reaction at the gas–solid interface allows dye desorption to be ignored. Therefore, numerous dyes are available for dye-sensitized SCR.


image file: d1ta02644e-f10.tif
Fig. 10 (a) Photo-SCR of NO over various metal oxide (1.0 wt%)-promoted TiO2 (GHSV: 50[thin space (1/6-em)]000 h−1). (b) Photo-SCR of NO over dye-modified TiO2 under visible light irradiation (dye loading: 12.5 μmol g−1, GHSV: 100[thin space (1/6-em)]000 h−1). (1) N3 dye, (2) Rose Bengal, (3) eosin Y, (4) Ru(bpy)3Cl2, (5) rhodamine B, (6) coumarin 343, (7) TCPP, (8) methylene blue, (9) Zn phthalocyanine, (10) Congo Red, (11) phthalocyanine, (12) RhCl3, (13) Indigo Carmine, (14) Cu phthalocyanine, and (15) carmine dyes. Reprinted with permission from ref. 109 and 121. Copyright© 2016 and 2015, Wiley.

Yao's group designed diverse types of photocatalytic systems for the photo-SCR of NO, mainly those relying on (i) cascadal electron transfer [LaFe1−xMnxO3/palygorskite,122 La1−xCexFeO3/palygorskite,123 LaFe1−xNixO3/palygorskite,124 La1−xPrxCoO3/palygorskite,125 CaTi1−xMnxO3−δ126], (ii) Z-scheme electron transfer [N-doped carbon quantum dots/PrFeO3,127 Fe2O3/SmFeO3/palygorskite,128 LaCoO3/palygorskite/reduced graphene oxide,129 Pr1−xCexFeO3/palygorskite,130 CeVO4/modified palygorskite131], and (iii) up-conversion (near-infrared light → UV and visible light) [CeO2/Pr3+/palygorskite132 and CeO2/palygorskite133] (Fig. 11 and Table 2). The metal ion content and hetero-element doping in mixed oxides altered the photocatalyst’s physical properties such as particle size, electronic band structure, surface acidity, and charge trapping sites, and the supports (palygorskite) were shown to prevent nanoparticle agglomeration and provide sites for NH3 adsorption. High NO conversion and the selective formation of N2 were achieved, and the mechanism of NO conversion to N2 was the same as that reported by Tanaka's group despite the difference in electron transfer pathways proposed. The Ag nano- and sub-nano-clusters incorporated in zeolites also promoted photo-assisted SCR under visible light irradiation (λ ≥ 390 nm), with activity determined by the reaction temperature (room temperature vs. 150 °C).134 Agnδ+ clusters were utilized as sensitizers because of their surface plasmon resonance and they favored the decomposition of NH2NO to N2 at 150 °C as opposed to the further oxidation of image file: d1ta02644e-t5.tif to NO and then to NO2 by singlet oxygen at room temperature.


image file: d1ta02644e-f11.tif
Fig. 11 Schematic diagrams of (a) cascadal electron transfer, (b) Z-scheme electron transfer, and (c) electron transfer in composite materials for the photocatalytic oxidation of NH3. Reprinted with permission from ref. 124 and 133. Copyright© 2018 and 2020, Elsevier. Reprinted with permission from ref. 127. Copyright© 2018, American Chemical Society.

DeNOxvia the oxidation of carbon-containing compounds. NH3 can be replaced by CH4, C3H8, C4H10, CO, CO(NH2)2, C2H5OH, carbon black, etc. for the photo-SCR of NO to N2 over TiO2-based catalysts under aerobic conditions. Bowering et al. tested the photocatalytic conversion of NO to N2 over TiO2 (P25) in the presence of CO as a reducing agent under UV light irradiation and showed that the reaction did not follow the Eley–Rideal mechanism but was rather driven by the adsorbed CO and NO (i.e., (i) COads + Oads → CO2(g); (ii) COads + 2NOads → N2Oads + CO2ads; (iii) COads + N2Oads → N2(g) + CO2ads).135 Therefore, the rate of NO conversion was the highest under CO-free conditions owing to the absence of competitive adsorption, whereas the highest selectivity for N2 was achieved in the presence of CO even though NO conversion was reduced. In the latter case, the surface hydroxyl groups significantly influenced selectivity (more hydroxyl groups led to better performance), and the loading of Ag on TiO2 markedly enhanced selectivity while reducing NO conversion, as Ag clusters acted as centers for electron–hole pair recombination.136 The electron defects (Ti3+, F+, and F centers) intentionally introduced on TiO2 by partial reduction could reduce NO under visible light irradiation, and the presence of CO as a regenerator of donor centers increased the selectivity for N2.137 Although extra experiments were also performed in the presence of hydrocarbons such as C2H6, C2H4, C3H8, propylene, C4H10, benzene, toluene, ethylbenzene, and o-xylene, the research goal was not the selective conversion of NO to N2 but the utilization of NO as an oxidant for the removal of VOC.138,139

Wu's group employed photo-SCR for denitrification in the presence of saturated hydrocarbons including CH4,140,141 C3H8,142–144 and C4H10[thin space (1/6-em)]145–147 as reducing agents, focusing on the synthesis of TiO2 and its structure/surface modification in Pd/TiO2, PtOxPdOy/TiO2, PdO/TiO2, Ag/TiO2, Cu/TiO2, Pt/TiO2, and TiO2 nanosheets. In the absence of co-catalysts on TiO2, the electron-donating behavior of hydrocarbons was not effectively utilized, unlike in the case of NH3. However, the temperature and the presence of moisture and oxygen were important for controlling NO conversion and selectivity for N2. For example, when PtOxPdOy/TiO2 was tested at temperatures of 25, 70, and 120 °C, the best performance was observed at the highest temperature when either oxygen or water vapor was present, as these conditions helped to avoid the accumulation of nitrate and the desorption of water vapor from active sites, respectively. On the other hand, the opposite trend was observed under vapor- and oxygen-free conditions because of the poor adsorption of C3H8 and NO (competitive adsorption as in the case of CO) on the catalyst surface at high temperature. Without PtOxPdOy catalysts, NO oxidation was dominant, and therefore, nitrate was formed as the major product, with its accumulation on the surface resulting in a decrease of activity with reaction time.

The utilization of urea, C2H5OH, and carbon black as reducing agents was also possible for the photocatalytic denitrification of NO to N2. In the case of TiO2 and urea co-supported carbon fiber, TiO2 and urea promoted the formation of NO2 and the sequential reduction of NO2 to N2 at room temperature, respectively, with moisture accelerating the desorption of NO2 from TiO2.148 In the case of Au/TiO2 + ethanol, the adsorption and dissociation of ethanol on Au particles or at the Au/TiO2 interface initiated the reaction at room temperature when C2H5Oads accepted a photogenerated electron. This reaction was also promoted by the by-products (CH3CHO, H2, CO, and CH4) formed by ethanol decomposition.149 Last, the reduction of NO to N2 was conducted over TiO2 along with the photocatalytic oxidation of carbon black to CO2 in the presence of O2 and moisture at 150 °C, providing the possibility of utilizing solid materials as reductants (Table 2).150

2.2.2. Photocatalytic removal of N2O(g) to N2. Several decades ago, the photocatalytic denitrification of N2O to N2 was investigated on ZnO at 371–431 °C under UV light irradiation, which induced the decomposition of N2O via a combination of thermocatalysis and photocatalysis.151 The thermocatalytic reaction followed first-order kinetics, while the photocatalytic reaction kinetics was more complicated and governed by N2O that was formed as an intermediate through electron transfer from ZnO to N2O. The mechanism proposed for n-type metal oxides in the dark at 20 °C comprises five steps and features step 2 as the rate-limiting step and Oads⋯MO+(s) as the dominant species because of the fast dissociation of N2Oads (Fig. 12a).152 Therefore, the N2 formation rate sharply increased and then saturated within the initial reaction stage over ZnO, in which case the kinetics was much faster than in the case of the photocatalytic reaction (Fig. 12b). Although the quantum efficiency of photo-assisted dissociation was small, an additional increase in N2 generation clearly appeared with time, and the reaction pathway (i.e., N2Oads + (e + h+) → image file: d1ta02644e-t6.tif → N2(g) + Oads) was specified by the migration of photogenerated charge carriers to the surface of ZnO. Electron paramagnetic resonance spectroscopy revealed that electrons are transferred to N2Oads (N2Oads + e → N2(g) + Oads), and Oads might be localized at oxide ion vacancies via migration, while holes can be trapped at oxide ions via migration through the lattice.153 Anpo et al. detected hyperfine splitting caused by one nitrogen atom at 77 K, which indicated the formation of either N2O or N2O2 (N2O + O → N2O2) on TiO2 supported by porous Vycor glass.154,155
image file: d1ta02644e-f12.tif
Fig. 12 (a) Mechanism of the conversion of N2O to N2 and O2 over a metal oxide surface. (b) Time-dependent production of N2 (VN) from N2O (10 Torr) over an activated ZnO surface at 20 °C under (i) dark and light on/off conditions indicated by arrows and (ii) continuous illumination. Reprinted with permission from ref. 152. Copyright© 1971, American Chemical Society. (c) Relative energy diagram for the photocatalytic decomposition of N2O on perfect anatase (001) facets. Reprinted with permission from ref. 173. Copyright© 2018, Royal Society of Chemistry. (d) Reaction diagram for the conversion of N2O to various adsorption and decomposition products (N2O*, image file: d1ta02644e-t13.tif, O*, image file: d1ta02644e-t14.tif, N2(g) and O2(g); * designates adsorbed species). (1–3) Dissociative adsorption of N2O on Ir(111), (4) formation of surface peroxides, and (5) O2 desorption from Ir(111). Reprinted with permission from ref. 177. Copyright© 2019, American Chemical Society.

Kudo et al. reported the denitrification of N2O over metal (Pt, Ag, and Cu)-loaded TiO2 in the presence of electron donors (water or/and methanol vapor) under UV light.156,157 Pt promoted the separation of electron–hole pairs, the dissociation of N2O, and the supply of adsorbed hydrogen atoms (i.e., H+ + e → H; N2O + 2H → N2 + H2O; N2O + H → N2 + OH), while water was oxidized on TiO2 (4OH + 4h+ → O2 + 2H2O). In the presence of both water and CH3OH vapor, the photocatalytic activity of Pt/TiO2 for N2O reduction was almost negligible, as the photogenerated electrons were selectively transferred to water to produce H2. Ag- and Cu-loaded TiO2 promoted denitrification, probably because of the facile dissociation of N2Oads on Au and Cu surfaces as well as the relatively high kinetic barrier for the reduction of water by electrons. Sano et al. further probed the removal of N2O over Ag/TiO2 in the presence of CH3OH vapor and showed that the photocatalytic performance was affected by the oxidation state of Ag.158 In particular, partially reduced Ag2O prepared by photodeposition was more active than metallic Ag, which was ascribed to Ag+-mediated charge transfer (N2Oads + e → N2Oads; N2Oads + Ag+→ N2 + Ag–O; 3Ag–O + CH3OHads + 3h+ → 3Ag+ + CO2 + 2H2O).

Metal ions (Cu+, Ag+, Pb2+, and Pr3+) were immobilized on the surface of metal oxides (SiO2, Al2O3, and SiO2/Al2O3) or incorporated inside ZSM-5 zeolite pores to promote the photo-assisted removal of N2O. Cu+-anchored metal oxides were prepared by an ion-exchange method with thermovacuum treatment, and linear two-coordinate and planar three-coordinate Cu+ ions were observed on SiO2/Al2O3 and Al2O3 or SiO2, respectively.159 In this case, Cu+ was assumed to undergo a (3d10),1S0 → (3d)9(4s)1,1D2 electronic transition under UV light irradiation, and electron transfer from the photo-excited Cu+ to N2O initiated denitrification. Photocatalytic activity was affected by the coordination geometry (linear or planar) and the aggregation state (isolated Cu+ monomer or Cu+–Cu+ dimer) of Cu+ and was the highest for the isolated linearly coordinated Cu+ monomer owing to the long lifetime of charge carriers coupled with the low accumulation of Oads due to O2 release. As Cu+ was incorporated in ZSM-5 and Y zeolite cavities, the type of Cu+ species depended on the degassing temperature, which suggested that the excited state of the Cu+–Cu+ dimer was an effective N2O quencher.160–162 In the case of Ag+-exchanged ZSM5, UV light irradiation induced the 4d10 → 4d95s1 transition of two-coordinate isolated Ag+ ions, and the complexation of Ag+ with N2O provided a channel for electron transfer from the excited Ag+ to the antibonding molecular orbital of N2O.163 For Pb2+-exchanged and Pr3+-supported catalysts, the reaction mechanisms were similar to those observed for Cu+- and Ag+-exchanged ones.164,165

Kočí's group reported diverse photocatalysts for the decomposition of N2O under UV light irradiation, e.g., ZnS/montmorillonite,166 cordierite/steatite/CeO2,167 TiO2/C3N4,168 WO3/C3N4,169 ZnO/C3N4,170 BiOIO3/C3N4,171 and BiVO4/C3N4.172 Among them, binary photocatalyst combinations helped to inhibit charge carrier recombination and therefore exhibited enhanced photocatalytic denitrification performances. Although some of these photocatalysts exhibited visible-light activity, all experiments were carried out under UV light. Moreover, the physicochemical interactions between the catalyst surface and N2O were not deeply investigated. To bridge this gap, Liu's group used DFT calculations to model the decomposition of N2O on TiO2,173 CeO2,174 BiVO4,175,176 BiMoO6,176 and Bi2WO6,176 obtaining results well correlated with experimental findings. For example, in the case of TiO2, the photogenerated electrons did not affect N2O adsorption, but the presence of oxygen vacancies or excited electrons promoted the N2O decomposition reaction. The surface-trapped electrons at five-coordinate Ti (Ti5c4+ + e → Ti5c3+) centers could act as active sites for N–O bond cleavage, with the reaction pathway depending on the adsorption geometry, i.e., on whether N2O (O[double bond, length as m-dash]N+[double bond, length as m-dash]NO–N+[triple bond, length as m-dash]N) was adsorbed on TiO2via the oxygen or the nitrogen end. In the case of decomposition on perfect anatase (001) facets, the N2O adsorbed on Ti3+via the oxygen end possessed an exothermic energy of 0.17 eV, and the O–N bond cleavage by the transfer of excited electrons from Ti3+ to N2O featured an exothermicity of 0.29 eV and produced N2 (Fig. 12c). On the other hand, the N2O adsorbed on Ti3+via the nitrogen end formed an intermediate bridging configuration (with a binding energy of 0.19 eV), and the N–O bond cleavage was characterized by an enthalpy change of −0.40 eV. Finally, N2 release from TiO2 was an endothermic (by 0.13 eV) process. The removal of O was ascribed to O discharge followed by recombination with another O atom, which proceeded via hole transfer to O and could decrease the energy barrier for O2 production.

An Al–Ir plasmonic antenna reactor combining plasmonic metallic antenna nanoparticles (Al nanocrystals) with nearby catalytic reactors (Ir nanoparticles) was designed for the photocatalytic conversion of N2O to N2 and O2.177 At high GHSVs (≥80[thin space (1/6-em)]000 h−1), the conversion efficiency reached 10%, and N2 and O2 were the only products formed. The apparent activation energy was maintained irrespective of illumination, which suggested that photothermal heating rather than hot carriers generated by the plasmon effect was responsible for N2O decomposition. As depicted in Fig. 12d, the pre-adsorption of N2O on Ir (step 1) and the dissociation of N2O (step 2) are not necessary because of the high exothermicity of the dissociative adsorption of N2O(g) into image file: d1ta02644e-t7.tif and O* at high operating temperatures (step 3). For fully saturated O*, the direct interaction between N2O(g) and O* can be driven by the Eley–Rideal mechanism to produce surface peroxide (step 4, moderately endothermic). Finally, the reaction is completed by the highly endothermic desorption of surface peroxide (image file: d1ta02644e-t8.tif; step 5), which was assumed to be the rate-limiting step for the overall N2O decomposition on Ir(111).

3. Oxidation of ammonia to N2 under aerobic and anaerobic conditions

Ammonia is one of the most valuable chemicals in agricultural and other industries, and has recently received much attention as a hydrogen carrier.178 As the manufacture of NH3 by the Haber–Bosch process is highly energy-intensive and consumes H2 that is mainly derived from fossil fuels, the economically feasible utilization of NH3 as a hydrogen carrier requires the development of highly active catalysts for the production of NH3 and its decomposition to H2 under mild conditions.179,180 Therefore, much effort has been directed at the establishment of new methods of ambient-condition N2 fixation, particularly those using renewable energy resources.181 Among these methods, the photocatalytic reduction of N2 to NH3 holds great promise, as the electrons and hydrogen are provided by sunlight and water, respectively, although the cleavage of the N[triple bond, length as m-dash]N bond in N2 at standard temperature and pressure is a big challenge because of the low solubility of this gas.182 From an environmental perspective, NH3 is not a useful chemical but a pollutant because of its high toxicity, corrosivity, odor, etc., and should therefore be effectively removed from air and water. The expansion of agricultural infrastructure to satisfy the increasing global food demand is facilitating the release of NH3 (from fertilizers, livestock excretions, etc.) to the atmosphere and water bodies.183 Most studies on photocatalysis target the oxidation of NH3 to N2 or NOx (2NH3 + 1.5O2 → N2 + 3H2O, ΔrG0298 = −652.41 kJ mol−1; 2NH3 + 4O2 → 2HNO3 + 2H2O, ΔrG0298 = −585.4 kJ mol−1), with comprehensive catalysts and relevant reaction mechanisms summarized in recent reviews.184,185 Herein, we briefly describe the photocatalytic decomposition of NH3 on TiO2 and present several examples of relatively high performance for the selective conversion of NH3 to N2 at room temperature.

Fig. 13a presents the mechanism of the photocatalytic oxidation of gas-phase NH3 on Pt/TiO2 in the presence/absence of water vapor under anaerobic conditions.186 Initially, NH3 is adsorbed on both Lewis- and Bronsted-acidic sites of TiO2 (mainly hydroxyl groups), and the reaction is initiated by the charge carriers generated under UV light irradiation. The electrons migrate to Pt nanoparticles to reduce protons and thus produce H2. The oxidation of adsorbed NH3 occurs via hole transfer, and the coupling of two amide radicals image file: d1ta02644e-t9.tif produces N2H4, which can be subsequently converted into H2 and N2H2. Finally, N2H2 self-decomposes into N2 and H2 or disproportionates into N2 and N2H4. As this process does not involve the formation of NOx, the H2[thin space (1/6-em)]:[thin space (1/6-em)]N2 molar ratio was recorded as 2.9, which was close to the theoretical value of 3.0 for the decomposition of NH3 to N2 and H2. Although the hole-mediated oxidation of image file: d1ta02644e-t10.tif to N through NH to release N2 is also possible, it is energetically unfavorable because of its higher net activation energy.160 Under dry conditions, the accumulation of NH4+ ions on TiO2 promotes catalyst deactivation, as these ions cannot easily migrate to Pt nanoparticles in the absence of water (Fig. 13b).


image file: d1ta02644e-f13.tif
Fig. 13 Proposed mechanism of the photocatalytic decomposition of NH3 on Pt/TiO2 in the (a) presence and (b) absence of water. Reprinted with permission from ref. 186. Copyright© 2012, American Chemical Society.

When TiO2 is used under aerobic and humid conditions, various nitrogen-containing species (e.g., NO, NO2, NO2, NO3, NO3, N2O, HONO, and N2H4) might be involved as intermediates or produced as by-products during NH3 oxidation,188–194 which complicates the selective production of N2. When the experiment was carried out by irradiating TiO2 in a flow tube with a stream of NH3-containing air, HONO was formed as an intermediate.190 The production of HONO was negligible in the absence of O2 and exhibited a volcano-type dependence on the concentration of NH3. The increase in [HONO] was ascribed to the photoreduction of NO2 (NH3 → NO2 → HONO), while the decrease in [HONO] at higher NH3 concentrations was ascribed to the saturation of surface-active sites according to the Langmuir–Hinshelwood model and the reaction with NH3 (NH3 + HONO → NH4NO2 → N2 + 2H2O). In a similar manner, [HONO] exhibited a volcano-like dependence on the relative humidity of the gas flow. Water accelerated the formation of HONO at low humidity, although excess water could occupy the pores of TiO2, hinder the access of NH3 to active sites, and facilitate the quenching of OH˙ to decrease [HONO]. In this experiment, NOx was formed as the major by-product. Instead of probing the complete conversion of NH3 to N2, almost all studies investigated the photocatalytic abatement of NH3 without analyzing the composition of the final products.

As mentioned earlier, the selective conversion of gaseous NH3 to N2 under aerobic conditions is challenging. From a practical viewpoint, operation under anaerobic conditions does not make sense, as the photocatalytic process is designed to remove few-ppm-level NH3 from air. In this regard, an anammox-like process aims to completely remove nitrogen species from aqueous systems (mainly NH3-containing wastewater) or use concentrated NH3 solutions as hydrogen carriers to provide H2 for fuel cells and should be more feasible owing to the ease of inert atmosphere generation via N2 or Ar purging. The protonation of NH3 (pKa ≈ 9.25) and the positive change of TiO2 surface charge (pHzpc 6–7 for P25) in acidic and neutral media cause electrostatic repulsion (i.e., NH4+ ↔ >Ti–OH2+), which hinders the adsorption of NH4+ and inhibits the photocatalytic reaction.195 Moreover, whereas NH4+ is stable against attack by OH˙, neutral NH3 is degraded by OH˙ under photocatalytic conditions.196–198 Therefore, high photocatalytic performance was achieved at pH 10–11, whereas an activity decrease was observed at higher pH, probably because of the low solubility of NH3 under these conditions.

The use of metal nanoparticles as co-catalysts offers a simple way to increase the yield and selectivity of photocatalytic processes, prolong charge carrier lifetime, and provide catalytically active sites. Among the various metal nanoparticles used in conjunction with TiO2, Pt nanoparticles exhibited an outstanding performance for the decomposition of NH3 into N2 and H2 under both oxic and anoxic conditions.199–201 Based on the calculated adsorption energies, Pt (−394 kJ mol−1) has a moderate atomic nitrogen affinity for N2 formation among the tested metals (e.g., Ag (−156 kJ mol−1), Au (−162 kJ mol−1), Rh (−448 kJ mol−1), Ru (−525 kJ mol−1)).202 In comparison with bare TiO2, which generated only nitrite and nitrate as end-products under air-saturated conditions, the loading of Pt (0.2 wt%) accelerated the reaction kinetics and promoted the evolution of N2 to reduce the total N content in the NH3 solution.170 Interestingly, the presence of O2 had little influence on the kinetics over Pt/TiO2, for which the efficiency of the NH3 to N2 conversion after 2 h irradiation equaled 65–70% in both air and N2. Pt nanoparticles on TiO2 probably stabilized NHx species generated as intermediates by OH˙-mediated chain reactions. When O2 was replaced by N2O, more OH radicals were formed through the reductive dissociation of N2O on Pt to increase the efficiency of the NH3 to N2 conversion to 80%. The photocatalytic conversion of NH3 and the selectivity for N2 simultaneously increased with the increase in the loading of Pt on TNTs under oxic conditions. In particular, an ammonia conversion of 100% (for [NH3]i = 20 ppm) and a selectivity of 87.5% were achieved after 3 h irradiation for Pt/TNTs (25 wt% Pt).200 Although the reductive dissociation of NH3 on Pt and the overoxidation of NH3 on TNTs might be responsible for the formation of N2 and NOx ions, respectively, it is still unclear whether the reductive dissociation of NH3 is energetically favorable or not, and the function of nitrogen hydrogen radicals on Pt as electron/hole recombination centers remains to be explored.

Under anaerobic conditions, the H2[thin space (1/6-em)]:[thin space (1/6-em)]N2 molar ratio achieved at alkaline pH using metallized photocatalysts (Pt/TiO2, Pt/Fe-doped TiO2, Ni/TiO2, Pt0.9Au0.1/TiO2, and Ru/ZnS) was close to the theoretical molar ratio (3[thin space (1/6-em)]:[thin space (1/6-em)]1). The main advantage of the anammox-like process is its ability to achieve both the complete removal of NH3 from wastewater and the recovery of H2 as a fuel for fuel cells at room temperature under sunlight. For example, in a highly concentrated solution (0.59 M), NH3 was decomposed at pH 10–12 over Pt/TiO2 (0.5 wt% Pt) to afford H2 and N2 in a 3[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio, and the catalyst performance was governed by Pt loading, pH, photocatalyst type, and co-catalyst type. Despite the lack of supporting evidence, Pt was assumed to provide active sites for the reduction of protons to H2, while the oxidation of NH3 occurred on TiO2. When Pt/Fe-doped TiO2 (0.5 wt% Pt and 1.0 wt% Fe) was tested in 0.59 M NH3 under UV light irradiation, a 3[thin space (1/6-em)]:[thin space (1/6-em)]1 (mol/mol) H2[thin space (1/6-em)]:[thin space (1/6-em)]N2 ratio was recorded.203 The higher H2 yield of Pt/Fe-doped TiO2 (27 μmol mgcat mol−1) than that of Pt/TiO2 (18 μmol mgcat mol−1) was due to the better absorption of visible light in the former case. Except for the case of Ni/TiO2, the loading of non-noble-metals (V, Cr, Mn, Fe, Co, and Cu) on TiO2 slightly decreased the H2 yield, whereas the H2[thin space (1/6-em)]:[thin space (1/6-em)]N2 molar ratio of 3[thin space (1/6-em)]:[thin space (1/6-em)]1 was maintained (0.59 M NH3).188 As seen in Fig. 14a, the amounts of N2 and H2 produced over Ni/TiO2 (0.5 wt% Ni) linearly increased with increasing irradiation time, and the reaction completely stopped in the dark. Isotope labeling experiments performed with D2O revealed that no D2 and HD were produced, i.e., the hydrogen in H2 stemmed from NH3 and not from water (Fig. 14b). This result indicates that the photodecomposition of NH3 occurred on the interface between metal nanoparticles and TiO2 and involved the direct migration of H˙ (formed by the hole-mediated reaction of NH3) to Pt. DFT calculations indicated the existence of two possible pathways for TiO2-based NH3 decomposition, namely (i) 2NH3, adsimage file: d1ta02644e-t11.tif + H2(g) → H2N–NH2 + H2(g) → ˙N[double bond, length as m-dash]N˙ + 3H2(g) → N2(g) + 3H2(g) and (ii) NH3, ads + NH3image file: d1ta02644e-t12.tif + H˙ + NH3 → NH2–NH3 + H˙ → H2N–NH2 + H2(g) → ˙N[double bond, length as m-dash]N˙ + 3H2(g) → N2(g) + 3H2(g). The formation of NH2–NH2 was probably assisted by metallic Ni. The loading of bimetallic alloy nanoparticles on TiO2 is also a good way to enhance the photocatalytic activity of monometallic nanoparticle/TiO2 hybrids, with the highest activity obtained for Pt0.9Au0.1 under UV light irradiation (Fig. 14c).204 As depicted in Fig. 14d, charge separation efficiency is determined by the Schottky barrier (φB; φB = metal work function (W) − electron affinity of the TiO2 conduction band (χ)). The introduction of Au into Pt reduces φB, which was calculated as 1.84, 1.62, and 0.97 eV for Pt/TiO2, Pt0.9Au0.1/TiO2, and Au/TiO2, respectively. Overly high and low φB values suppress electron separation and promote reverse electron transfer, thus decelerating photocatalytic reactions. The decomposition of NH3 into N2 and H2 was also carried out using other photocatalysts such as RuO2–NiO–SrTiO3,205 ZnO,206 and Ru/ZnS,207 the activities of which were much lower than that of Pt/TiO2. The visible light-induced decomposition of NH3 into N2 and H2 was also attempted in a dye-sensitized system comprising a homogeneous tris(bipyridine)ruthenium(II) (Ru(bpy)32+) dye, methyl viologen as an electron mediator, and O2 as an electron acceptor.208 Under visible light irradiation, Ru(bpy)33+ oxidized NH3 and was converted to the original state, Ru(bpy)32+.


image file: d1ta02644e-f14.tif
Fig. 14 (a) Time profiles of H2 and N2 production by the photodecomposition of NH3 over Ni/TiO2 (1.0 wt%) in the dark and under illumination. (b) Time profiles of gas-phase product yields for the photocatalytic decomposition of NH3 over Ni/TiO2 in D2O. Reprinted with permission from ref. 187. Copyright© 2017, Elsevier. (c) Amounts of H2 and N2 evolved during the 6 h photocatalytic decomposition of NH3 on Pt0.9M0.1/TiO2 (M: Au, Pd, Cu, Ni, and Ag, total metal loading on TiO2: 0.1 mol%). (d) Schematic electronic structure of a metal/semiconductor interface. Evac, EF, W, φB, and χ denote the vacuum level, Fermi energy level, metal work function, Schottky barrier, and the electron affinity of the semiconductor conduction band, respectively. Reprinted with permission from ref. 204. Copyright© 2020, American Chemical Society.

4. Photoelectrochemical denitrification and ammonia oxidation

Only a few studies deal with photoelectrochemical denitrification and ammonia oxidation, focusing on the recovery of N2 from nitrogen species. The photoelectrochemical denitrification of nitrite (1 mM NaNO2 at pH 7) was first achieved in 1999 using a three-electrode system (counter electrode (CE) = Pt wire, reference electrode (RE) = saturated calomel electrode (SCE), and working electrode (WE) = roughened Ag electrode) in 0.1 M Na2SO4 as an electrolyte under laser irradiation (362, 413, 457, 476, 488, 496, 514, and 647 nm) and a nitrogen atmosphere.209 Denitrification was initiated by the excitation of Ag via plasmon resonance and the electrochemical current generated at −1.0 VSCE. Notably, irradiation brought about not only an increase in cathodic current but also a positive shift of the onset potential. The quantum efficiency was estimated as 0.04% without the analysis of real-time NO2, NO3, N2, and NH3 concentrations. Nitrate reduction was believed to involve the electrochemical nitrate to nitrite conversion followed by the photoelectrochemical reduction of nitrite to NH3 and N2. The photoelectrochemical nitrate to nitrite conversion was also observed for Ag nanopyramids (1 M NaNO3 at pH 5.7 under Ar), in which case the plasmon resonance of Ag resulted in an almost 100% faradaic efficiency at −1.0 VRHE.210

The photoelectrochemical nitrate to nitrite reduction was also promoted by semiconducting photocathode materials such as p-GaInP2, nanoporous p-Si, and CuI/PbI2. In the case of p-GaInP2, data were collected in a three-electrode system (CE = Pt black, RE = Ag/AgCl, and WE = p-GaInP2) in 0.1 M HNO3 + 0.5 M NH4NO3 as an electrolyte (pH 1) under simulated solar light at an air mass (AM) of 1.5 G.211 The faradaic efficiency of nitrate reduction was calculated as 80%, and the incident-photon-to-current efficiency (IPCE) at −1.0 VAg/AgCl was recorded as 100, 60, and 5% under excitation at 400, 580, and 610 nm, respectively. As a close to zero current was obtained in the dark, illumination was concluded to stimulate the rate-limiting step, and the catalytically active sites were assumed to be Ga and/or In. For nanoporous p-Si under similar conditions, the faradaic efficiency of nitrate reduction at −0.6 VAg/AgCl equaled 65%, and no NH3 and N2 were observed.212 In the case of CuI–PbI2, the faradaic efficiency of nitrate reduction in 0.1 M NaNO3 exceeded 52%, and the IPCE at 400 nm was around 15%.213 The bubbles evolved on the photoelectrode surface probably contained N2 rather than H2, as no H2 signal was observed by gas chromatography. Interestingly, isotope labeling experiments performed in Ar-saturated 0.1 M Na15NO3 solution (98% 15N) revealed that the generation of NH3 was due to an external contamination and not nitrate reduction.

n-type semiconductors can be used as photoanodes for the water oxidation-induced conversion of nitrate to N2. In the presence of NH3 as an electron donor (i.e., using the same concept as that discussed in Section 2.2.1, the photo-SCR deNOx), ammonia oxidation and denitrification simultaneously occurred over TiO2 and Pt black, respectively, in the absence of a bias voltage under UV light irradiation (1 mM NH3, 100 mM KNO3) (Fig. 15a).214 When a mixture of pig urine/wash water (1/4) containing NH4+, NO3, and NO2 was tested under aerobic conditions, the following concentration decreases were observed after 24 h: NH4+ (2580 → 166 ppm), NO3 (18.6 → 17.0 ppm), and NO2 (4.84 → 3.17 ppm). The imbalance in the removal of NOx and NH4+ was ascribed to the competitive reduction of oxygen to H2O. Similarly, in a biophotochemical cell, H2O or biorefractory organics were oxidized at the photoanode, while denitrification proceeded at the biocathode.215,216 The biocathode was prepared using activated sludge as an inoculum and was separated from the TiO2 photoanode by a cation exchange membrane. As seen in Fig. 15b, the concentration of nitrate continuously decreased under illumination, whereas the abiotic cathode did not show any activity. Indeed, NO2 and N2O were formed as intermediates, but the concentration of these intermediates and NO3 decreased to zero after 30 h (Fig. 15c). NH4+ ions were always present at levels below the detection limit, which indicated that nitrate was selectively converted to N2. The faradaic efficiency of the cathode was estimated as 97%, and a small number of electrons was assumed to be consumed by microbial growth.


image file: d1ta02644e-f15.tif
Fig. 15 (a) Photoelectrochemical denitrification of NO3 to N2 over a TiO2 photoanode connected to a Pt cathode in the presence of NH3 and H2O as electron donors under Ar. Reprinted with permission from ref. 214. Copyright© 2009, Royal Society of Chemistry. (b) Decrease of nitrate level under a photo-generated current during on–off intermittent illumination and (c) change of nitrogen oxide levels with time. Reprinted with permission from ref. 215. Copyright© 2017, American Chemical Society. (d) Proposed mechanism for the oxidation of NH3 over a CuO/Co3O4 photocathode in the presence of peroxydisulfate. Reprinted with permission from ref. 219. Copyright© 2020, Elsevier.

The photoelectrochemical oxidation of ammonia to N2 can be accomplished using the anodic or cathodic reaction to control photogenerated holes or radical species (hydroxyl or sulfate radicals) activated by electrons, respectively.217–220 In the employed system (CE = Pt wire, RE = Ag/AgCl, and WE = TiO2 photoanode in 10 M NH3 + 0.1 M KNO3 at pH 14.1), the H2[thin space (1/6-em)]:[thin space (1/6-em)]N2 molar ratio equaled 3.08 under short-circuit conditions after 2 h irradiation.188 The holes in TiO2 oxidized NH3 to N2, and the electrons transferred to Pt reduced water to H2. The OH˙ and SO4˙ species generated by the activation of peroxydisulfate (S2O82− + e → SO4˙ + SO42− and SO4˙ + H2O → H+ + OH˙ + SO42−) at the CuO/Co3O4 photocathode oxidized NH3 to N2 (Fig. 15d).219 The removal of 96.1% NH3 (100 ppm) was achieved under visible light irradiation, and the reactive sites were identified as Co and Cu species.

5. Summary and outlook

There are various methods of decreasing the levels of reactive nitrogen compounds, which can include systematic crop rotation, optimization of the timing and amount of fertilizer input, the breeding or development of genetically engineered varieties of crops for increased nitrogen utilization efficiency, direct up-cycling of used nitrogen to microbial protein, and the development of artificial denitrification/ammonia oxidation processes powered by renewable energy. Among them, the solar-powered photocatalytic and photoelectrochemical approaches are promising and future-oriented ways to treat aqueous and airborne NOx, N2O, and NH3 because of their economically feasible and environmentally benign nature. However, the low efficiency and selectivity to N2 and the scale-up problems are still a bottleneck for practical applications. Basically, the efficiency of artificial solar-powered denitrification/ammonia oxidation can be determined by (i) the absorbance of photocatalysts, (ii) the electronic structure of semiconducting materials, (iii) the recombination of charge carriers (i.e., the lifetime of photogenerated electrons and holes), (iv) the control of surface properties to suppress undesired reactions such as ammonification and the re-oxidation of intermediates and products, and (v) the surroundings (e.g., the presence/absence of oxygen, different types of electron donors, pH, humidity, etc.). Due to the different reaction pathways, the development of new photocatalysts and the systematic design of reactors have to be different depending on the treatment of aqueous and gas phase NOx, N2O, and NH3. In an aqueous system, the bimetallic nanoparticles loaded on TiO2 (e.g., Pd–Cu and Pt–Cu) showed a high performance of NO3 conversion selective to N2, which gives a hint that the separation of the catalytic sites, the reduction of NO3 to NO2 and the further reduction of NO2 to N2, is important to increase the selectivity. Precise control of the N[thin space (1/6-em)]:[thin space (1/6-em)]H ratio at surface active sites of catalysts including Lewis acid sites and defect sites is particularly required to suppress ammonification and thus selectively convert NO2 to harmless N2. The reason why it is hard to reach a high conversion efficiency of NO3 or NO2 as well as a high selectivity to N2 is due to the competitive charge transfer to H+, O2, and H2O and the re-oxidation of intermediates and by-products. In the case of the photocatalytic treatment of airborne NOx and N2O to N2, it has been thoroughly investigated during the last three decades, offering the advantages of using water instead of the explosive H2 and toxic NH3, operation at standard temperature and pressure, a net zero carbon emission in the case of operation under natural sunlight, and the availability of cheap and environmentally benign materials. Although various strategies such as structure and morphology control, co-catalyst loading, heteroatom doping, and hybridization with different types of materials have been developed to overcome the bottleneck of the conversion of NOx and N2O to N2, the problems of insufficiently selective charge transfer, low solar light absorption, poor catalytic activity, need for noble metals as co-catalysts, and the lack of long-term durability need to be addressed to minimize the environmental impact of airborne NOx and N2O. In order to increase the possibility of commercialization, the addition of sacrificial hole scavengers should be necessary, where organic pollutants, particularly persistent organic pollutants (POPs), and ammonia (or carbon monoxide) are good candidates for wastewater and polluted air treatment, respectively, to compensate the operation cost.

Although ammonia is a very important feedstock, its high toxicity, corrosivity, and noxious odor make it a pollutant from the perspective of the environmental management of the nitrogen cycle. Given the large annual production of ammonia via the Haber–Bosch process and the low nitrogen use efficiency of ammonia-based fertilizers, ammonia should be effectively removed from air and water on a comparable scale. Recently, considering ammonia as a hydrogen carrier, the development of highly active catalysts for the decomposition of ammonia to H2 and N2 under mild conditions is highly desired. Although the conventional photocatalytic processes have focused on ammonia abatement, directing the production of nitrate instead of N2, with a future-oriented point of view, the photocatalytic recovery of H2 from concentrated ammonia solution seems quite promising for fuel cell applications. The oxidation of adsorbed NH3 occurring via photogenerated holes does not involve the formation of NOx, but the H2[thin space (1/6-em)]:[thin space (1/6-em)]N2 molar ratio becomes close to the theoretical value of 3[thin space (1/6-em)]:[thin space (1/6-em)]1 through the decomposition of NH3 to N2 and H2, which is driven at standard temperature and pressure under illumination.

A photoelectrochemical cell can selectively control the reduction and oxidation reaction of nitrogen-containing species, in which the photogenerated electron–hole pairs are easily separated and consequently participate in the denitrification and anammox upon applying extra bias. To date, very few studies have been reported, in particular targeting the removal of toxic nitrite and ammonia from wastewater; however, the application should be more suitable for anammox in order to secure H2 from concentrated ammonia. Contrary to photocatalysis, it does not need to separate N2 and H2 because the oxidation and the reduction are proceeded in the anode and cathode, respectively, which is comparted by a membrane. Indeed, the addition of an electrolyte is unnecessary in that the pH of concentrated ammonia solution (>12) is conductive enough to transport ions in the electrolyte. The in-depth investigation and successful development of photoelectrochemical ammonia oxidation systems will enable a counterpart of (photo)electrocatalytic nitrogen fixation, in other words the combination of the production and the utilization of ammonia as a hydrogen carrier.

It is time to take this issue seriously and think about it, and photocatalysis is the greenest way to restore the nitrogen cycle with a future-oriented technology. In order to go one step further to commercialization, the following can be considered: (i) the development of new materials to overcome the intrinsic problems of photocatalysts, (ii) the control of composition, morphology, and size of catalysts (e.g., high entropy alloy, single atom catalyst, etc.), (iii) the systematic modification of photocatalysts including hybridization such as ternary and quaternary composites, co-doping, anchoring homogeneous sensitizers or promoters, selective surface passivation, etc., (iv) the separation of catalytic sites by the control of the boundary between the catalysts and supporter or by a Janus structure, (v) the precise control of the micro-environment on the catalysts or electrodes, (vi) the finding of suitable POPs and greenhouse gases that cannot be removed by conventional treatment or typical AOPs, (vii) the design of a photo-reactor and its scale-up, (viii) the combination with other processes such as the pretreatment or final treatment through biological processes, and (ix) in situ analysis (e.g., time-resolved surface enhanced infrared absorption/Raman spectroscopy) to unveil the real-time charge transfer and the formation of intermediates for the optimization of desired reactions.

Author contributions

Cheolwoo Park: investigation, visualization, writing – original draft. Hyelim Kwak: investigation, visualization, writing – original draft. Gun-hee Moon: conceptualization, supervision, writing – original draft. Wooyul Kim: conceptualization, supervision, writing – review & editing.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This research was financially supported by the Basic Science Research Program (NRF-2019R1C1C1006833) funded by the Korea government (MSIT) through the National Research Foundation of Korea (NRF), the Ecological Imitation-Based Environmental Pollution Management Technology Development Project funded by the Korea government (MOE) through KEITI (No. 2019002790008), the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (NRF-2020M3H4A3106354), and the KIST internal project (3E311191) funded by the Korea Institute of Science and Technology (KIST).

Notes and references

  1. A. Bernhard, Nat. Educ. Knowl., 2010, 2, 1–8 Search PubMed.
  2. D. E. Canfield, A. N. Glazer and P. G. Falkowski, Science, 2010, 330, 192–196 CrossRef CAS PubMed.
  3. J. N. Galloway, A. M. Leach, A. Bleeker and J. W. Erisman, Philos. Trans. R. Soc., B, 2013, 368, 20130120 CrossRef PubMed.
  4. M. A. Sutton, C. M. Howard, T. K. Adhya, E. Baker, J. Baron, A. Basir, W. Brownlie, C. Cordovil, W. de Vries, V. Eory, R. Green, H. Harmens, K. W. Hicks, R. Jeffery, D. Kanter, L. Lassaletta, A. Leip, C. Masso, T. H. Misselbrook, E. Nemitz, S. P. Nissanka, O. Oenema, S. Patra, M. Pradhan, J. Ometto, R. Purvaja, N. Raghuram, R. Ramesh, N. Read, D. S. Reay, E. Rowe, A. SanzCobena, S. Sharma, K. R. Sharp, U. Skiba, J. U. Smith, I. van der Beck, M. Vieno, and H. J. M. van Grinsven, Nitrogen - Grasping the Challenge. A Manifesto for Science-In-Action through the International Nitrogen Management System. Summary Report, Ecology & Hydrology, Edinburgh, UK, 2019 Search PubMed.
  5. Food and Agriculture Organization, FAO Statistical Databases, 2006, Rome, available at http://faostat.fao.org/default.aspx Search PubMed.
  6. International Energy Agency, IEA Statistical Databases, 2015, Paris, available at http://iea.org/data-and-statistics Search PubMed.
  7. Revised 1996 IPCC Guidelines for National Greenhouse Gas Inventories, ed. J. T. Houghton, L. G. Meria Filho, K. Lim, I. Trennton, I. Mamaty, Y. Bonduki, D. J. Griggs and B. A. Callander, IPCC, OECD, IEA, 1996 Search PubMed.
  8. D. Laffoley and J. M. Baxter, Ocean Deoxygenation: Everyone’s Problem, IUCN, Global Marine and Polar Programme, 2019 Search PubMed.
  9. N. Gruber and J. N. Galloway, Nature, 2008, 451, 293–296 CrossRef CAS PubMed.
  10. S. Matassa, D. J. Batstone, T. Hülsen, J. Schnoor and W. Verstraete, Environ. Sci. Technol., 2015, 49, 5247–5254 CrossRef CAS PubMed.
  11. S. P. Seitzinger, C. Kroeze, A. F. Bouwman, N. Caraco, F. Dentener and R. V. Styles, Estuaries, 2002, 25, 640–655 CrossRef CAS.
  12. M. B. Peoples, J. Brockwell, D. F. Herridge, I. J. Rochester, B. J. R. Alves, S. Urquiaga, R. M. Boddey, F. D. Dakora, S. Bhattarai, S. L. Maskey, C. Sampet, B. Rerkasem, D. F. Khan, H. Hauggaard-Nielsen and E. S. Jensen, Symbiosis, 2009, 48, 1–17 CrossRef CAS.
  13. Y. Lan, J. Chen, H. Zhang, W.-X. Zhang and J. Yang, J. Mater. Chem. A, 2020, 8, 15853–15863 RSC.
  14. W. Hong, L. Su, J. Wang, M. Jiang, Y. Ma and J. Yang, Chem. Commun., 2020, 56, 14685–14688 RSC.
  15. F. Ni, Y. Ma, J. Chen, W. Luo and J. Yanga, Chin. Chem. Lett., 2021, 32, 2073–2078 CrossRef CAS.
  16. L. Su, D. Han, G. Zhu, H. Xu, W. Luo, L. Wang, W. Jiang, A. Dong and J. Yang, Nano Lett., 2019, 19, 5423–5430 CrossRef CAS PubMed.
  17. H. Xu, J. Wu, W. Luo, Q. Li, W. Zhang and J. Yang, Small, 2020, 16, 2001775 CrossRef CAS PubMed.
  18. D. R. Keeney, R. L. Chen and D. A. Graetz, Nature, 1971, 233, 66–67 CrossRef CAS PubMed.
  19. A. Kapoor and T. Viraraghavan, J. Environ. Eng., 1997, 123, 371–380 CrossRef CAS.
  20. H. O. N. Tugaoen, S. Garcia-Segura, K. Hristovski and P. Westerhoff, Sci. Total Environ., 2017, 599–600, 1524–1551 CrossRef PubMed.
  21. Z. Geng, Z. Chen, Z. Li, X. Qi, X. Yang, W. Fan, Y. Guo, L. Zhang and M. Huo, Dalton Trans., 2018, 47, 11104–11112 RSC.
  22. H. Xu, Y. Li, M. Ding, W. Chen, K. Wang and C. Lu, ACS Sustainable Chem. Eng., 2018, 6, 7042–7051 CrossRef CAS.
  23. S. Roy, J. Phys. Chem. C, 2020, 124, 28345–28358 CrossRef CAS.
  24. H. Zhang, Z. Liu, Y. Li, C. Zhang, Y. Wang, W. Zhang, L. Wang, L. Niu, P. Wang and C. Wang, Appl. Surf. Sci., 2020, 503, 144092 CrossRef CAS.
  25. J. E. Silveira, A. R. Ribeiro, J. Carbajo, G. Pliego, J. A. Zazo and J. A. Casas, Water Res., 2021, 200, 117250 CrossRef CAS PubMed.
  26. L. Wang, W. Fu, Y. Zhuge, J. Wang, F. Yao, W. Zhong and X. Ge, Chemosphere, 2021, 278, 130298 CrossRef CAS PubMed.
  27. T. Caswell, M. W. Dlamini, P. J. Miedziak, S. Pattisson, P. R. Davies, S. H. Taylor and G. J. Hutchings, Catal. Sci. Technol., 2020, 10, 2082–2091 RSC.
  28. H. Kominami, K. Kitsui, Y. Ishiyama and K. Hashimoto, RSC Adv., 2014, 4, 51576–51579 RSC.
  29. H. Gekko, K. Hashimoto and H. Kominami, Phys. Chem. Chem. Phys., 2012, 14, 7965–7970 RSC.
  30. W. L. Silver, D. J. Herman and M. K. Firestone, Ecology, 2001, 82, 2410–2416 CrossRef.
  31. E. M. Sander, B. Virdis and S. Freguia, RSC Adv., 2015, 5, 86572–86577 RSC.
  32. H. Hirakawa, M. Hashimoto, Y. Shiraishi and T. Hirai, ACS Catal., 2017, 7, 3713–3720 CrossRef CAS.
  33. H. Kominami, A. Furusho, S.-y. Murakami, H. Inoue, Y. Kera and B. Ohtani, Catal. Lett., 2001, 76, 31–34 CrossRef CAS.
  34. K. Akihiko, D. Kazunari, M. Ken-ichi and O. Takaharu, Chem. Lett., 1987, 16, 1019–1022 CrossRef.
  35. M. Duca and M. T. M. Koper, Energy Environ. Sci., 2012, 5, 9726–9742 RSC.
  36. Z. Hou, J. Chu, C. Liu, J. Wang, A. Li, T. Lin and C. P. François-Xavier, Chem. Eng. J., 2021, 415, 2082–2091 CrossRef.
  37. S. Lee, S. Kim, C. Park, W. Kim, S. Ryu and W. Choi, Energy Environ. Sci., 2021 10.1039/D1EE01342D.
  38. P. Li, Z. Jin, Z. Fanga and G. Yu, Energy Environ. Sci., 2021, 14, 3522–3531 RSC.
  39. M. Shand and J. A. Anderson, Catal. Sci. Technol., 2013, 3, 879–899 RSC.
  40. H.-i. Kim, K. Kim, S. Park, W. Kim, S. Kim and J. Kim, Sep. Purif. Technol., 2019, 209, 580–587 CrossRef CAS.
  41. L. Lei, W. Wang, C. Wang, H. Fan, A. K. Yadav, N. Hu, Q. Zhong and P. Müller-Buschbaum, J. Mater. Chem. A, 2020, 8, 23812–23819 RSC.
  42. Q. Liu, S. Wang, Q. Ren, T. Li, G. Tu, S. Zhong, Y. Zhao and S. Bai, J. Mater. Chem. A, 2021, 9, 1552–1562 RSC.
  43. X. Li, H. Jiang, C. Ma, Z. Zhu, X. Song, X. Li, H. Wang, P. Huo and X. Chen, J. Mater. Chem. A, 2020, 8, 18707–18714 RSC.
  44. S. Lee, S. Kim, C. Park, G.-h. Moon, H.-J. Son, J.-O. Baeg, W. Kim and W. Choi, ACS Sustainable Chem. Eng., 2020, 8, 3709–3717 CrossRef CAS.
  45. H. Lee, S. Jee, R. Kim, H.-T. Bui, B. Kim, J.-K. Kim, K. Park, W. Choi, W. Kim and K. Choi, Energy Environ. Sci., 2020, 13, 519–526 RSC.
  46. E.-T. Yun, H.-Y. Yoo, W. Kim, H.-E. Kim, G. Kang, H. Lee, S. Lee, T. Park, C. Lee, J.-H. Kim and J. Lee, Appl. Catal., B, 2017, 203, 475–484 CrossRef CAS.
  47. Y. Zhang, W. Hu, D. Wang, B. J. Reinhart and J. Huang, J. Mater. Chem. A, 2021, 9, 6180–6187 RSC.
  48. S. Weon, E. Choi, H. Kim, J. Kim, H.-J. Park, S.-m. Kim, W. Kim and W. Choi, Environ. Sci. Technol., 2018, 52, 9330–9340 CrossRef CAS PubMed.
  49. Q. Guo, F. Liang, Z. Sun, Y. Wang, X.-B. Li, S.-G. Xia, Z. C. Zhang, L. Huang and L.-Z. Wu, J. Mater. Chem. A, 2020, 8, 22601–22606 RSC.
  50. J. Sá, C. A. Agüera, S. Gross and J. A. Anderson, Appl. Catal., B, 2009, 85, 192–200 CrossRef.
  51. J. A. Anderson, Catal. Today, 2011, 175, 316–321 CrossRef CAS.
  52. W. Gao, R. Jin, J. Chen, X. Guan, H. Zeng, F. Zhang and N. Guan, Catal. Today, 2004, 90, 331–336 CrossRef CAS.
  53. S. Rengaraj and X. Z. Li, Enhanced photocatalytic reduction reaction over Bi3+-TiO2 nanoparticles in presence of formic acid as a hole scavenger, Chemosphere, 2007, 66, 930–9383 CrossRef CAS PubMed.
  54. D. D. B. Luiz, S. L. F. Andersen, C. Berger, H. J. José and R. D. F. P. M. Moreira, J. Photochem. Photobiol. A, 2012, 246, 36–44 CrossRef CAS.
  55. J. R. Pan, C. Huang, W. Hsieh and B. Wu, Sep. Purif. Technol., 2012, 84, 52–55 CrossRef CAS.
  56. K. Kobwittaya and S. Sirivithayapakorn, APCBEE Proc., 2014, 10, 321–325 CrossRef CAS.
  57. S.-E. Bae, K. L. Stewart and A. A. Gewirth, J. Am. Chem. Soc., 2007, 129, 10171–10180 CrossRef CAS PubMed.
  58. I. Sanjuán, L. García-Cruz, J. Solla-Gullón, E. Expósito and V. Montiel, Electrochim. Acta, 2020, 340, 135914 CrossRef.
  59. S. Hamid, M. A. Kumar, J.-I. Han, H. Kim and W. Lee, Green Chem., 2017, 19, 853–866 RSC.
  60. H. Shin, S. Jung, S. Bae, W. Lee and H. Kim, Environ. Sci. Technol., 2014, 48, 12768–12774 CrossRef CAS PubMed.
  61. T. Gu, W. Teng, N. Bai, Z. Chen, J. Fan, W.-x. Zhang and D. Zhao, J. Mater. Chem. A, 2020, 8, 9545–9553 RSC.
  62. S. Jung, S. Bae and W. Lee, Environ. Sci. Technol., 2014, 48, 9651–9658 CrossRef CAS PubMed.
  63. A. J. Calandra, C. Tamayo, J. Herrera and A. J. Arvía, Electrochim. Acta, 1972, 17, 2035–2053 CrossRef CAS.
  64. L. Li, Z. Xu, F. Liu, Y. Shao, J. Wang, H. Wan and S. Zheng, J. Photochem. Photobiol., A, 2010, 212, 113–121 CrossRef CAS.
  65. H. Kominami, T. Nakaseko, Y. Shimada, A. Furusho, H. Inoue, S.-y. Murakami, Y. Kera and B. Ohtani, Chem. Commun., 2005, 2933–2935 RSC.
  66. O. S. G. P. Soares, M. F. R. Pereira, J. J. M. Órfão, J. L. Faria and C. G. Silva, Chem. Eng. J., 2014, 251, 123–130 CrossRef CAS.
  67. Z. Hou, F. Chen, J. Wang, C. P. François-Xavier and T. Wintgens, Appl. Catal., B, 2018, 232, 124–134 CrossRef CAS.
  68. J. A. Zazo, P. García-Muñoz, G. Pliego, J. E. Silveira, P. Jaffe and J. A. Casas, Appl. Catal., B, 2020, 273, 118930 CrossRef CAS.
  69. H. Kato and A. Kudo, Phys. Chem. Chem. Phys., 2002, 4, 2833–2838 RSC.
  70. L. Mohapatra and K. Parida, J. Mater. Chem. A, 2016, 4, 10744–10766 RSC.
  71. N. Dewangan, W. M. Hui, S. Jayaprakash, A.-R. Bawah, A. J. Poerjoto, T. Jie, A. Jangam, K. Hidajat, S. Kawi, Catal. Today, 356, 490–513 Search PubMed.
  72. S. Zhang, Y. Zhao, R. Shi, C. Zhou, G. I. N. Waterhouse, L.-Z. Wu, C.-H. Tung and T. Zhang, Adv. Energy Mater., 2020, 10, 1901973 CrossRef CAS.
  73. D. P. Sahoo, K. K. Das, S. Patnaik and K. Parida, Inorg. Chem. Front., 2020, 7, 3695–3717 RSC.
  74. Y. Zhao, G. Chen, T. Bian, C. Zhou, G. I. N. Waterhouse, L.-Z. Wu, C.-H. Tung, L. J. Smith, D. O'Hare and T. Zhang, Adv. Mater., 2015, 27, 7824–7831 CrossRef CAS PubMed.
  75. M. Adachi and A. Kudo, Chem. Lett., 2012, 41, 1007–1008 CrossRef CAS.
  76. X. Li, S. Wang, H. An, G. Dong, J. Feng, T. Wei, Y. Ren and J. Ma, Appl. Surf. Sci., 2021, 539, 148257 CrossRef CAS.
  77. M. Yue, R. Wang, B. Ma, R. Cong, W. Gao and T. Yang, Catal. Sci. Technol., 2016, 6, 8300–8308 RSC.
  78. R. Wang, M. Yue, R. Cong, W. Gao and T. Yang, J. Alloys Compd., 2015, 651, 731–736 CrossRef CAS.
  79. B. Wang, B. An, Z. Su, L. Li and Y. Liu, Chemosphere, 2021, 269, 128754 CrossRef CAS PubMed.
  80. G. Liu, S. You, M. Ma, H. Huang and N. Ren, Environ. Sci. Technol., 2016, 50, 11218–11225 CrossRef CAS PubMed.
  81. H.-T. Ren, S.-Y. Jia, J.-J. Zou, S.-H. Wu and X. Han, Appl. Catal., B, 2015, 176–177, 53–61 CrossRef CAS.
  82. Y. Li and F. Wasgestian, J. Photochem. Photobiol., A, 1998, 112, 255–259 CrossRef CAS.
  83. J. A. Anderson, Catal. Today, 2012, 181, 171–176 CrossRef CAS.
  84. B. Bems, F. C. Jentoft and R. Schlögl, Appl. Catal., B, 1999, 20, 155–163 CrossRef CAS.
  85. F. Zhang, Y. Pi, J. Cui, Y. Yang, X. Zhang and N. Guan, J. Phys. Chem. C, 2007, 111, 3756–3761 CrossRef CAS.
  86. J. M. A. Freire, M. A. F. Matos, D. S. Abreu, H. Becker, I. C. N. Diógenes, A. Valentini and E. Longhinotti, J. Environ. Chem. Eng., 2020, 8, 103844 CrossRef CAS.
  87. A. Oita, A. Malik, K. Kanemoto, A. Geschke, S. Nishijima and M. Lenzen, Nat. Geosci., 2016, 9, 111–115 CrossRef CAS.
  88. D. S. Reay, E. A. Davidson, K. A. Smith, P. Smith, J. M. Melillo, F. Dentener and P. J. Crutzen, Nat. Clim. Change, 2012, 2, 410–416 CrossRef CAS.
  89. G. Busca, L. Lietti, G. Ramis and F. Berti, Appl. Catal., B, 1998, 18, 1–36 CrossRef CAS.
  90. L. Yang, A. Hakki, F. Wang and D. E. Macphee, Appl. Catal., B, 2018, 222, 200–208 CrossRef CAS.
  91. Z. Gu, B. Zhang, Y. Asakura, S. Tsukuda, H. Kato, M. Kakihana and S. Yin, Appl. Surf. Sci., 2020, 521, 146213 CrossRef CAS.
  92. V. Khanal, N. O. Balayeva, C. Günnemann, Z. Mamiyev, R. Dillert, D. W. Bahnemann and V. R. Subramanian, Appl. Catal., B, 2021, 291, 119974 CrossRef CAS.
  93. G. Liu, H. Xia, W. Zhang, L. Song, Q. Chen and Y. Niu, J. Hazard. Mater., 2021, 418, 126337 CrossRef CAS PubMed.
  94. J. V. C. d. Carmo, C. L. Lima, G. Mota, A. M. S. Santos, L. N. Costa, A. Ghosh, B. C. Viana, M. Silva, J. M. Soares, S. Tehuacanero-Cuapa, R. Lang, A. C. Oliveira, E. Rodríguez-Castellón and E. Rodríguez-Aguado, Materials, 2021, 14, 2181 CrossRef PubMed.
  95. W. Kim, T. Tachikawa, G.-h. Moon, T. Majima and W. Choi, Angew. Chem., Int. Ed., 2014, 53, 14036–14041 CrossRef CAS PubMed.
  96. Q. Wu and R. van de Krol, J. Am. Chem. Soc., 2012, 134, 9369–9375 CrossRef CAS PubMed.
  97. Q. Wu, Q. Zheng and R. van de Krol, J. Phys. Chem. C, 2012, 116, 7219–7226 CrossRef CAS.
  98. G. Dong, D. L. Jacobs, L. Zang and C. Wang, Appl. Catal., B, 2017, 218, 515–524 CrossRef CAS.
  99. M. Anpo, Y. Shioya, H. Yamashita, E. Giamello, C. Morterra, M. Che, H. H. Patterson, S. Webber and S. Ouellette, J. Phys. Chem., 1994, 98, 5744–5750 CrossRef CAS.
  100. M. Matsuoka, E. Matsuda, K. Tsuji, H. Yamashita and M. Anpo, J. Mol. Catal. A: Chem., 1996, 107, 399–403 CrossRef CAS.
  101. W.-S. Ju, M. Matsuoka and M. Anpo, Catal. Lett., 2001, 71, 91–93 CrossRef CAS.
  102. M. Anpo, S. G. Zhang, H. Mishima, M. Matsuoka and H. Yamashita, Catal. Today, 1997, 39, 159–168 CrossRef CAS.
  103. Y. Hu, G. Martra, J. Zhang, S. Higashimoto, S. Coluccia and M. Anpo, J. Phys. Chem. B, 2006, 110, 1680–1685 CrossRef CAS PubMed.
  104. M. Matsuoka and M. Anpo, in Catalysis by Unique Metal Ion Structures in Solid Matrices, ed. G. Centi, B. Wichterlová and A. T. Bell, Springer, Dordrecht, 2001, ch. Photocatalysis of Cations Incorporated within Zeolites, vol. 13, pp. 249–262 Search PubMed.
  105. M. Anpo, M. Matsuoka, K. Hanou, H. Mishima, H. Yamashita and H. H. Patterson, Coord. Chem. Rev., 1998, 171, 175–184 CrossRef CAS.
  106. M. Anpo and J. M. Thomas, Chem. Commun., 2006, 3273–3278 RSC.
  107. N. W. Cant and J. R. Cole, J. Catal., 1992, 134, 317–330 CrossRef CAS.
  108. T. Shishido, K. Teramura and T. Tanaka, Catal. Sci. Technol., 2011, 1, 541–551 RSC.
  109. A. Yamamoto, K. Teramura and T. Tanaka, Chem. Rec., 2016, 16, 2268–2277 CrossRef CAS PubMed.
  110. K. Teramura, T. Tanaka and T. Funabiki, Langmuir, 2003, 19, 1209–1214 CrossRef CAS.
  111. K. Teramura, T. Tanaka, S. Yamazoe, K. Arakaki and T. Funabiki, Appl. Catal., B, 2004, 53, 29–36 CrossRef CAS.
  112. S. Yamazoe, T. Okumura, K. Teramura and T. Tanaka, Catal. Today, 2006, 111, 266–270 CrossRef CAS.
  113. Y. Ji and Y. Luo, J. Phys. Chem. C, 2014, 118, 6359–6364 CrossRef CAS.
  114. S. Yamazoe, Y. Masutani, T. Shishido and T. Tanaka, Res. Chem. Intermed., 2008, 34, 487–494 CrossRef CAS.
  115. S. Yamazoe, Y. Masutani, K. Teramura, Y. Hitomi, T. Shishido and T. Tanaka, Appl. Catal., B, 2008, 83, 123–130 CrossRef CAS.
  116. R. Jin, Z. Wu, Y. Liu, B. Jiang and H. Wang, J. Hazard. Mater., 2009, 161, 42–48 CrossRef CAS PubMed.
  117. Y.-C. Chou and Y. Ku, Chem. Eng. J., 2013, 225, 734–743 CrossRef CAS.
  118. S. Yamazoe, K. Teramura, Y. Hitomi, T. Shishido and T. Tanaka, J. Phys. Chem. C, 2007, 111, 14189–14197 CrossRef CAS.
  119. G. Zhang, G. Kim and W. Choi, Energy Environ. Sci., 2014, 7, 954–966 RSC.
  120. A. Yamamoto, Y. Mizuno, K. Teramura, S. Hosokawa, T. Shishido and T. Tanaka, Catal. Sci. Technol., 2015, 5, 556–561 RSC.
  121. A. Yamamoto, K. Teramura, S. Hosokawa, T. Shishido and T. Tanaka, ChemCatChem, 2015, 7, 1818–1825 CrossRef CAS.
  122. X. Li, X. Yan, S. Zuo, X. Lu, S. Luo, Z. Li, C. Yao and C. Ni, Chem. Eng. J., 2017, 320, 211–221 CrossRef CAS.
  123. H. Zhang, X. Li, Y. Hui, L. Yu, Q. Xia, S. Luo and C. Yao, J. Mater. Sci.: Mater. Electron., 2017, 28, 9371–9377 CrossRef CAS.
  124. X. Li, H. Shi, W. Zhu, S. Zuo, X. Lu, S. Luo, Z. Li, C. Yao and Y. Chen, Appl. Catal., B, 2018, 231, 92–100 CrossRef CAS.
  125. K. Wei, X. Yan, S. Zuo, W. Zhu, F. Wu, X. Li, C. Yao and X. Liu, Clays Clay Miner., 2019, 67, 348–356 CrossRef CAS.
  126. Z. Zhang, H. Lü, X. Li, X. Li, S. Ran, Z. Chen, Y. Yang, X. Wu and L. Li, ACS Sustainable Chem. Eng., 2019, 7, 10299–10309 CrossRef CAS.
  127. X. Li, H. Shi, X. Yan, S. Zuo, Y. Zhang, T. Wang, S. Luo, C. Yao and C. Ni, ACS Sustainable Chem. Eng., 2018, 6, 10616–10627 CrossRef CAS.
  128. X. Li, H. Shi, T. Wang, Y. Zhang, X. Lu, S. Zuo, Z. Li and C. Yao, J. Taiwan Inst. Chem. Eng., 2018, 89, 119–128 CrossRef CAS.
  129. X. Li, X. Yan, X. Lu, S. Zuo, Z. Li, C. Yao and C. Ni, J. Catal., 2018, 357, 59–68 CrossRef.
  130. X. Li, H. Shi, X. Yan, S. Zuo, Y. Zhang, Q. Chen, C. Yao and C. Ni, J. Catal., 2019, 369, 190–200 CrossRef CAS.
  131. X. Li, Z. Wang, H. Shi, D. Dai, S. Zuo, C. Yao and C. Ni, J. Hazard. Mater., 2020, 386, 121977 CrossRef CAS PubMed.
  132. X. Li, H. Zhang, H. Lü, S. Zuo, Y. Zhang and C. Yao, Environ. Sci. Pollut. Res., 2019, 26, 12842–12850 CrossRef CAS PubMed.
  133. X. Li, Z. Wang, X. Chu, B. Gao, S. Zuo, W. Liu and C. Yao, Appl. Clay Sci., 2020, 199, 105871 CrossRef CAS.
  134. H. I. Hamoud, M. Lafjah, F. Douma, O. I. Lebedev, F. Djafri, V. Valchev, M. Daturi and M. El-Roz, Sol. Energy, 2019, 189, 244–253 CrossRef CAS.
  135. N. Bowering, G. S. Walker and P. G. Harrison, Appl. Catal., B, 2006, 62, 208–216 CrossRef CAS.
  136. N. Bowering, D. Croston, P. G. Harrison and G. S. Walker, Int. J. Photoenergy, 2007, 2007, 090752 CrossRef.
  137. A. A. Lisachenko, R. V. Mikhailov, L. L. Basov, B. N. Shelimov and M. Che, J. Phys. Chem. C, 2007, 111, 14440–14447 CrossRef CAS.
  138. C. H. Ao, S. C. Lee, C. L. Mak and L. Y. Chan, Appl. Catal., B, 2003, 42, 119–129 CrossRef CAS.
  139. S. Poulston, M. V. Twigg and A. P. Walker, Appl. Catal., B, 2009, 89, 335–341 CrossRef CAS.
  140. Y.-T. Wu, Y.-H. Yu, V.-H. Nguyen and J. C. S. Wu, Res. Chem. Intermed., 2015, 41, 2153–2164 CrossRef CAS.
  141. Y.-H. Yu, I.-H. Su and J. C. S. Wu, Environ. Technol., 2010, 31, 1449–1458 CrossRef CAS PubMed.
  142. J. Lasek, Y.-H. Yu and J. C. S. Wu, Environ. Technol., 2012, 33, 2133–2141 CrossRef CAS PubMed.
  143. Y.-H. Yu, Y.-T. Pan, Y.-T. Wu, J. Lasek and J. C. S. Wu, Catal. Today, 2011, 174, 141–147 CrossRef CAS.
  144. I. H. Su and J. C. S. Wu, Catal. Commun., 2009, 10, 1534–1537 CrossRef CAS.
  145. J. C.-C. Yu, V.-H. Nguyen, J. Lasek, S.-W. Chiang, D. X. Li and J. C. S. Wu, Appl. Catal., A, 2016, 523, 294–303 CrossRef CAS.
  146. J. C.-C. Yu, V.-H. Nguyen, J. Lasek and J. C. S. Wu, Appl. Catal., B, 2017, 219, 391–400 CrossRef CAS.
  147. J. C.-C. Yu, V.-H. Nguyen, J. Lasek, D. X. Li and J. C. S. Wu, Catal. Commun., 2016, 84, 40–43 CrossRef CAS.
  148. J. Miyawaki, T. Shimohara, N. Shirahama, A. Yasutake, M. Yoshikawa, I. Mochida and S.-H. Yoon, Appl. Catal., B, 2011, 110, 273–278 CrossRef CAS.
  149. G. Halasi, T. Bánsági and F. Solymosi, J. Catal., 2015, 325, 60–67 CrossRef CAS.
  150. L. Liao, S. Heylen, S. P. Sree, B. Vallaey, M. Keulemans, S. Lenaerts, M. B. J. Roeffaers and J. A. Martens, Appl. Catal., B, 2017, 202, 381–387 CrossRef CAS.
  151. G. Blyholder and K. Tanaka, J. Phys. Chem., 1971, 75, 1037–1043 CrossRef CAS.
  152. J. Cunningham, J. J. Kelly and A. L. Penny, J. Phys. Chem., 1971, 75, 617–625 CrossRef CAS.
  153. N. B. Wong, Y. B. Taarit and J. H. Lunsford, J. Chem. Phys., 1974, 60, 2148–2151 CrossRef CAS.
  154. M. Anpo, N. Aikawa and Y. Kubokawa, J. Chem. Soc., Chem. Commun., 1984, 644–645 RSC.
  155. M. Anpo, N. Aikawa, Y. Kubokawa, M. Che, C. Louis and E. Giamello, J. Phys. Chem., 1985, 89, 5689–5694 CrossRef CAS.
  156. A. Kudo and T. Sakata, Chem. Lett., 1992, 21, 2381–2384 CrossRef.
  157. A. Kudo and H. Nagayoshi, Catal. Lett., 1998, 52, 109–111 CrossRef CAS.
  158. T. Sano, N. Negishi, D. Mas and K. Takeuchi, J. Catal., 2000, 194, 71–79 CrossRef CAS.
  159. M. Matsuoka, W.-S. Ju, K. Takahashi, H. Yamashita and M. Anpo, J. Phys. Chem. B, 2000, 104, 4911–4915 CrossRef CAS.
  160. K. Ebitani, M. Morokuma, J.-H. Kim and A. Morikawa, J. Catal., 1993, 141, 725–728 CrossRef CAS.
  161. K. Ebitani, M. Morokuma, J.-H. Kim and A. Morikawa, J. Chem. Soc., Faraday Trans., 1994, 90, 377–381 RSC.
  162. K. Ebitani, M. Morokuma and A. Morikawa, in Stud. Surf. Sci. Catal., ed. J. Weitkamp, H. G. Karge, H. Pfeifer and W. Hölderich, Elsevier, 1994, vol. 84, pp. 1501–1506 Search PubMed.
  163. W.-S. Ju, M. Matsuoka, K. Iino, H. Yamashita and M. Anpo, J. Phys. Chem. B, 2004, 108, 2128–2133 CrossRef CAS.
  164. W.-S. Ju, M. Matsuoka, H. Yamashita and M. Anpo, J. Synchrotron Radiat., 2001, 8, 608–609 CrossRef CAS PubMed.
  165. K. Ebitani, Y. Hirano and A. Morikawa, J. Catal., 1995, 157, 262–265 CrossRef CAS.
  166. L. Obalová, M. Šihor, P. Praus, M. Reli and K. Kočí, Catal. Today, 2014, 230, 61–66 CrossRef.
  167. M. Valášková, K. Kočí and J. Kupková, Microporous Mesoporous Mater., 2015, 207, 120–125 CrossRef.
  168. K. Kočí, M. Reli, I. Troppová, M. Šihor, J. Kupková, P. Kustrowski and P. Praus, Appl. Surf. Sci., 2017, 396, 1685–1695 CrossRef.
  169. M. Reli, L. Svoboda, M. Šihor, I. Troppová, J. Pavlovský, P. Praus and K. Kočí, Environ. Sci. Pollut. Res. Int., 2018, 25, 34839–34850 CrossRef CAS PubMed.
  170. K. Kočí, M. Reli, I. Troppová, M. Šihor, T. Bajcarová, M. Ritz, J. Pavlovský and P. Praus, Catalysts, 2019, 9, 735 CrossRef.
  171. V. Matějka, M. Šihor, M. Reli, A. Martaus, K. Kočí, M. Kormunda and P. Praus, Mater. Sci. Semicond. Process., 2019, 100, 113–122 CrossRef.
  172. P. Praus, J. Lang, A. Martaus, L. Svoboda, V. Matějka, M. Kormunda, M. Šihor, M. Reli and K. Kočí, J. Inorg. Organomet. Polym., 2019, 29, 1219–1234 CrossRef CAS.
  173. L. Wang, W. Song, J. Deng, H. Zheng, J. Liu, Z. Zhao, M. Gao and Y. Wei, Nanoscale, 2018, 10, 6024–6038 RSC.
  174. W. Song, L. Wang, Y. Gao, J. Deng, M. Jing, H. Zheng, J. Liu, Z. Zhao, M. Gao and Y. Wei, J. Mater. Chem. A, 2018, 6, 19241–19255 RSC.
  175. L. Wang, J. Liu, W. Song, H. Wang, Y. Li, J. Liu, Z. Zhao, J. Tan, Z. Duan and J. Deng, Chem. Eng. J., 2019, 366, 504–513 CrossRef CAS.
  176. J. Liu, L. Wang, W. Song, M. Zhao, J. Liu, H. Wang, Z. Zhao, C. Xu and Z. Duan, ACS Sustainable Chem. Eng., 2019, 7, 2811–2820 CrossRef CAS.
  177. D. F. Swearer, H. Robatjazi, J. M. P. Martirez, M. Zhang, L. Zhou, E. A. Carter, P. Nordlander and N. J. Halas, ACS Nano, 2019, 13, 8076–8086 CrossRef CAS PubMed.
  178. A. Klerke, C. H. Christensen, J. K. Nørskov and T. Vegge, J. Mater. Chem., 2008, 18, 2304–2310 RSC.
  179. Q. Wang, J. Guo and P. Chen, J. Energy Chem., 2019, 36, 25–36 CrossRef.
  180. F. Schüth, R. Palkovits, R. Schlögl and D. S. Su, Energy Environ. Sci., 2012, 5, 6278–6289 RSC.
  181. X. Cui, C. Tang and Q. Zhang, Adv. Energy Mater., 2018, 8, 1800369 CrossRef.
  182. Q. Han, H. Jiao, L. Xiong and J. Tang, Adv. Mater., 2021, 2, 564–581 RSC.
  183. E. Stokstad, Science, 2014, 343, 238 CrossRef CAS PubMed.
  184. K. Vikrant, K.-H. Kim, F. Dong and D. A. Giannakoudakis, ACS Catal., 2020, 10, 8683–8716 CrossRef CAS.
  185. S. Zhang, Z. He, X. Li, J. Zhang, Q. Zang and S. Wang, Nanoscale Adv., 2020, 2, 3610–3623 RSC.
  186. H. Yuzawa, T. Mori, H. Itoh and H. Yoshida, J. Phys. Chem. C, 2012, 116, 4126–4136 CrossRef CAS.
  187. A. Utsunomiya, A. Okemoto, Y. Nishino, K. Kitagawa, H. Kobayashi, K. Taniya, Y. Ichihashi and S. Nishiyama, Appl. Catal., B, 2017, 206, 378–383 CrossRef CAS.
  188. A. C. Sola, D. Garzón Sousa, J. Araña, O. González Díaz, J. M. Doña Rodríguez, P. Ramírez de la Piscina and N. Homs, Catal. Today, 2016, 266, 53–61 CrossRef CAS.
  189. H. Mozzanega, J. M. Herrmann and P. Pichat, J. Phys. Chem., 1979, 83, 2251–2255 CrossRef CAS.
  190. M. A. Kebede, N. K. Scharko, L. E. Appelt and J. D. Raff, J. Phys. Chem. Lett., 2013, 4, 2618–2623 CrossRef CAS.
  191. B. Boulinguiez, A. Bouzaza, S. Merabet and D. Wolbert, J. Photochem. Photobiol., A, 2008, 200, 254–261 CrossRef CAS.
  192. M. Chen, J. Ma, B. Zhang, F. Wang, Y. Li, C. Zhang and H. He, Appl. Catal., B, 2018, 223, 209–215 CrossRef CAS.
  193. G. Zhang, J. Ruan and T. Du, ACS ES&T Eng., 2021, 1, 310–325 Search PubMed.
  194. J. Feng, X. Zhang, G. Zhang, J. Li, W. Song and Z. Xu, Chemosphere, 2021, 274, 129689 CrossRef CAS PubMed.
  195. X. Zhu, S. R. Castleberry, M. A. Nanny and E. C. Butler, Environ. Sci. Technol., 2005, 39, 3784–3791 CrossRef CAS PubMed.
  196. R. Zellner and I. W. M. Smith, Chem. Phys. Lett., 1974, 26, 72–74 CrossRef CAS.
  197. E.-M. Bonsen, S. Schroeter, H. Jacobs and J. C. Broekaert, Chemosphere, 1997, 35, 1431–1445 CrossRef CAS.
  198. S. Kim and W. Choi, Environ. Sci. Technol., 2002, 36, 2019–2025 CrossRef CAS PubMed.
  199. J. Lee, H. Park and W. Choi, Environ. Sci. Technol., 2002, 36, 5462–5468 CrossRef CAS PubMed.
  200. H.-H. Ou, M. R. Hoffmann, C.-H. Liao, J.-H. Hong and S.-L. Lo, Appl. Catal., B, 2010, 99, 74–80 CrossRef CAS.
  201. H. Kominami, H. Nishimune, Y. Ohta, Y. Arakawa and T. Inaba, Appl. Catal., B, 2012, 111–112, 297–302 CrossRef CAS.
  202. A. C. A. de Vooys, M. T. M. Koper, R. A. van Santen and J. A. R. van Veen, J. Electroanal. Chem., 2001, 506, 127–137 CrossRef CAS.
  203. J. Nemoto, N. Gokan, H. Ueno and M. Kaneko, J. Photochem. Photobiol., A, 2007, 185, 295–300 CrossRef CAS.
  204. Y. Shiraishi, S. Toi, S. Ichikawa and T. Hirai, ACS Appl. Nano Mater., 2020, 3, 1612–1620 CrossRef CAS.
  205. L. Qing-shui, D. Kazunari, N. Shuichi, O. Takaharu and T. Kenzi, Chem. Lett., 1983, 12, 321–324 CrossRef.
  206. M. Reli, M. Edelmannová, M. Šihor, P. Praus, L. Svoboda, K. K. Mamulová, H. Otoupalíková, L. Čapek, A. Hospodková, L. Obalová and K. Kočí, Int. J. Hydrogen Energy, 2015, 40, 8530–8538 CrossRef CAS.
  207. A. Iwase, K. Ii and A. Kudo, Chem. Commun., 2018, 54, 6117–6119 RSC.
  208. M. Kaneko, N. Katakura, C. Harada, Y. Takei and M. Hoshino, Chem. Commun., 2005, 3436–3438 RSC.
  209. J. Zheng, T. Lu, T. M. Cotton and G. Chumanov, J. Phys. Chem. B, 1999, 103, 6567–6572 CrossRef CAS.
  210. Y. Kim, E. B. Creel, E. R. Corson, B. D. McCloskey, J. J. Urban and R. Kostecki, Adv. Energy Mater., 2018, 8, 1800363 CrossRef.
  211. H. Wang and J. A. Turner, Energy Environ. Sci., 2013, 6, 1802–1805 RSC.
  212. M. Kan, D. Yue, J. Jia and Y. Zhao, Electrochim. Acta, 2015, 177, 366–369 CrossRef CAS.
  213. E. Kecsenovity, S. T. Kochuveedu, J.-P. Chou, D. Lukács, Á. Gali and C. Janáky, Sol. RRL, 2021, 5, 2000418 CrossRef CAS.
  214. R. Saito, H. Ueno, J. Nemoto, Y. Fujii, A. Izuoka and M. Kaneko, Chem. Commun., 2009, 3231–3233 RSC.
  215. H.-Y. Cheng, X.-D. Tian, C.-H. Li, S.-S. Wang, S.-G. Su, H.-C. Wang, B. Zhang, H. M. A. Sharif and A.-J. Wang, Environ. Sci. Technol., 2017, 51, 12948–12955 CrossRef CAS PubMed.
  216. S. Su, Y. Zhang, W. Hu, X. Zhang, D. Ju, C. Jia and J. Liu, Bioelectrochemistry, 2020, 132, 107439 CrossRef CAS PubMed.
  217. M. Kaneko, N. Gokan, N. Katakura, Y. Takei and M. Hoshino, Chem. Commun., 2005, 1625–1627 RSC.
  218. X. Fan, Y. Zhou, G. Zhang, T. Liu and W. Dong, Appl. Catal., B, 2019, 244, 396–406 CrossRef CAS.
  219. C. Yan and L. Liu, J. Hazard. Mater., 2020, 388, 121793 CrossRef CAS PubMed.
  220. Y. Qu, X. Song, X. Chen, X. Fan and G. Zhang, Chem. Eng. J., 2020, 382, 123048 CrossRef CAS.

Footnote

These authors contributed equally.

This journal is © The Royal Society of Chemistry 2021