Next Article in Journal
Recent Progress of Printing Technologies for High-Efficient Organic Solar Cells
Previous Article in Journal
Effect of Metal Atom in Zeolitic Imidazolate Frameworks (ZIF-8 & 67) for Removal of Dyes and Antibiotics from Wastewater: A Review
Previous Article in Special Issue
A Study on the Evaluation Methods of Nitrogen Oxide Removal Performance of Photocatalytic Concrete for Outdoor Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Challenges of Integrating the Principles of Green Chemistry and Green Engineering to Heterogeneous Photocatalysis to Treat Water and Produce Green H2

by
Fernanda Anaya-Rodríguez
,
Juan C. Durán-Álvarez
*,
K. T. Drisya
and
Rodolfo Zanella
*
Instituto de Ciencias Aplicadas y Tecnología, Universidad Nacional Autónoma de México, Circuito Exterior S/N, Coyoacán, Ciudad de Mexico 04510, Mexico
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(1), 154; https://doi.org/10.3390/catal13010154
Submission received: 30 November 2022 / Revised: 31 December 2022 / Accepted: 4 January 2023 / Published: 9 January 2023

Abstract

:
Nowadays, heterogeneous photocatalysis for water treatment and hydrogen production are topics gaining interest for scientists and developers from different areas, such as environmental technology and material science. Most of the efforts and resources are devoted to the development of new photocatalyst materials, while the modeling and development of reaction systems allowing for upscaling the process to pilot or industrial scale are scarce. In this work, we present what is known on the upscaling of heterogeneous photocatalysis to purify water and to produce green H2. The types of reactors successfully used in water treatment plants are presented as study cases. The challenges of upscaling the photocatalysis process to produce green H2 are explored from the perspectives of (a) the adaptation of photoreactors, (b) the competitiveness of the process, and (c) safety. Throughout the text, Green Chemistry and Engineering Principles are described and discussed on how they are currently being applied to the heterogeneous photocatalysis process along with the challenges that are ahead. Lastly, the role of automation and high-throughput methods in the upscaling following the Green Principles is discussed.

1. Introduction

There are two milestones in heterogeneous photocatalysis [1]. First, the group of Muller successfully degraded isopropanol using ZnO under UV irradiation in 1969. Later, in 1972, the group of Fujishima and Honda were able to produce hydrogen via the water splitting reaction by irradiating TiO2 with UV light. From 1975 to 1985, the first generation of photocatalysts (most of them TiO2-based) was developed for H2 production via the water splitting reaction, aiming at studying the semiconductor/solution interface under UV irradiation and looking for new and more efficient semiconductors. During the second-generation of photocatalysts (1986–2000), the investigation on modified semiconductors for a better response under visible light irradiation began [2], which was continued in the third generation that started in 2001 [3]. The corpus of literature was focused on modifying the photocatalysts by doping or coupling with other semiconductors to form heterostructures, including the S- and Z-schemes [4,5]. In environmental engineering, heterogeneous photocatalysis has been identified as a promising alternative for water purification and H2 production [6,7,8]. Upon the irradiation of the semiconductor with light of proper energy, the electrons at the highest occupied molecular orbital (HOMO, the valence band) are excited and thus promoted to the lowest unoccupied molecular orbital (LUMO, the conduction band), resulting in the generation of the charge carriers known as the hole/electron pair (Equation (1)). Then, the charge carriers migrate to the surface of the crystalline semiconductor (Figure 1). Once there, the photo-holes (h+) react with the adsorbed molecules, producing oxidative species. As shown in Equations (2) and (3), water molecules are oxidized to produce OH species that react with organic pollutants resulting in intermediates and, at the best conditions, reaching the mineralization. In some cases, photo-holes can directly react with the adsorbed organic molecules (Equation (4)), although it depends on the nucleophilic nature of the reagents (e.g., the occurrence of aromatic rings or resonance structures within the molecule). The reactive oxygen species are able not only to degrade recalcitrant organic pollutants but also kill pathogens, oxidize heavy metals, and destroy antibiotic resistance genes (Figure 1) [9].
Semiconductor + hv     h + +   e
h + +   H 2 O     OH   +   H +
OH +   Organic   molecules     Intermediantes     CO 2 + H 2 O
h + +   Organic   molecules     Intermediates
The photo-electrons in the conduction band react with the dissolved oxygen adsorbed on the surface of the photocatalyst, producing the superoxide species (O2), which are capable to oxidize organic pollutants (Equations (5)–(6)). Additionally, photoelectrons can directly react with pollutants to generate intermediates (Equation (7)); this process will depend on the electrophilic nature of the organic molecules (e.g., sulfonic moieties). Under reductive conditions, the photoelectrons are able to produce H2 through the water splitting reaction (Equation (8) and Figure 1).
e   +   O 2     O 2
  O 2   +   Organic   molecules     Intermediates     CO 2 + H 2 O
e +   Organic   molecules     Intermediates
e   + 2 H 2 O     2 H + + 2 OH   H 2
To achieve this, the potential of the conduction band must be higher than that necessary for the H+ reduction (i.e., 0 eV vs. NHE at pH = 0), plus an overpotential [10]. This condition constrains the number of semiconductors that can be used for H2 production, with nitrides, halides, and calcogenides (e.g., metal sulfides) as the most commonly chosen materials [3,11]. In the photocatalytic water splitting reaction, the photo-holes in the valence band must be scavenged by nucleophilic molecules, known as sacrificial agents, in order to avoid the re-oxidation of the produced hydrogen. The most commonly used sacrificial agents are light alcohols and carboxylic acids, sulfates, and other inorganic ions [10].
Overall, the photocatalytic process comprises three fundamental steps: (a) the absorption of light by the semiconductor producing the charge carriers; (b) the migration of the charge carriers to the surface of the crystal; and (c) the reaction of the reactive oxygen species and/or charge carriers with the adsorbed species to degrade organic pollutants and produce hydrogen (Figure 1). These steps are significantly affected by intrinsic material conditions, such as crystallinity, particle size, surface area, superficial charge, and optoelectronic properties, as well as the characteristics of the reaction medium, including pH, light penetration, composition, etc. Detailed information on these topics can be found in previous extensive reviews [12,13].
For benchmarking purposes, the International Organization for Standardization (ISO) has established five types of tests to assess the photocatalytic activity of ceramic materials, which are briefly described in Table 1.
Through these tests, it is possible to compare the activity of photocatalysts under standardized conditions. However, some parts are severely outdated, since using the degradation of azo dyes is not completely appropriate to measure the photocatalytic efficiency, as this type of organic molecules can sensitize the semiconductor, resulting in an experimental artifact. To date, the list of standard tests does not include photocatalytic H2 generation, although some documents (ISO 22734) suggest that some related tests are nearly ready.
Hydrogen is considered as a clean energy source due to its superior energy conversion compared to natural gas, while its combustion produces zero emissions of greenhouse gases [14]. Most of the commercially available H2 is produced from different feedstocks, such as electricity, biomass, and fossil fuels [15,16]. It has made room for the development of cleaner processes to produce green H2, with lower emissions and reduced energy inputs by harnessing the renewable sources and thus generating multiple desired outputs [14]. In this sense, producing green H2 through the photocatalytic water splitting reaction is a cleaner process compared to other sources [17]. Moreover, wastewater can be valorized as a raw material for hydrogen production. This can be achieved either through the aqueous reforming process that converts light hydrocarbons and water into CO and H2 at low temperatures [18,19] or by the heterogeneous photocatalysis process. For the latter, the dissolved organic matter, especially the light compounds, are used as sacrificial agents that impede the recombination of the photo-produced charge carriers at the time photo-electrons reduce the water molecule [20]. Even when the efficiency of the H2 production is low, the conjunctive action of the oxidation—reduction reactions to purify water while producing green H2 results in significant valorization of wastewater that is worth further investigation.
Currently, H2 is mainly produced from fossil fuels and the cost is from 2 to 4 times less than H2 production through the water splitting process [15]. For example, the costs of producing hydrogen from conventional fossil fuels varies between 1.34–2.27 USD kg−1, while using biomass as raw material costs between 1.77 and 2.05 USD kg−1 [21]. Conversely, the levelized cost of photocatalytic H2 production was estimated as 18.32 USD kg−1 in 2021 (making some important assumptions, such as the remarkedly high performance of the photocatalysts) [22]. To date, the efficiency of the photocatalytic water splitting reaction is below 1% (measured as the sun to hydrogen conversion, STH). Remarkable investigations are advancing to surpass this threshold and achieve more than 5% of STH conversion and, under this scenario, the cost in the production of green H2 would decrease to 1.6–3.5 USD kg−1 [23].
Current research has been focused on the synthesis of highly efficient photocatalysts to remove organic pollutants from water and to produce green H2 using low energy or renewable light sources, such as sunlight. We consider it is time to migrate from laboratory tests to upscaled systems, which include reaction systems testing real water and wastewater and microplant, pilot-plant, or industrial-scaled systems. As shown in Figure 2, through the last decade, most of the research has been centered on the synthesis of new photocatalysts, while the development of new photoreactors and the upscaling of the heterogeneous photocatalysis process have been left out.
Indeed, the discovery of new and highly effective photocatalysts through optimum synthetic routes and using high-throughput methods is a fertile field with new materials awaiting to be discovered. Nevertheless, the upscaling of photocatalytic reactors from the laboratory and pilot-scale is essential to globally implement these technologies, especially in regions with high sunlight irradiation. Mild and eco-friendly conditions must be prioritized to obtain sustainable upscaled systems, integrating Green Engineering and Green Chemistry Principles (GEP and GCP, respectively) to address environmental pollution and energy demand from a Life Cycle Analysis perspective [24], as these principles can be applied from the synthesis steps to the applications. Chemicals are unavoidable to achieve economic and social development, but their nature and mode of production can be tuned in accordance to the Green Principles, paving the way for sustainable development and aiming at improving the productivity of the processes by minimizing their harm [25].
The main objective of this review is to explore the current state of the art on the upscaling of the heterogeneous photocatalysis process, as well as to identify the integration of the Green Principles for the discovery and synthesis of new photocatalysts and the upscaling itself. The currently identified and foreseen challenges in the upscaling of green heterogeneous photocatalysis process are described and briefly discussed.

2. Integrating the Green Principles to the Heterogeneous Photocatalytic Process

The concept of green chemistry was introduced by the United States Pollution Prevention Act (1990) and the Green Chemistry Principles (GCP) were presented by Paul Anastas and John Warner in their search for a safer synthesis method (Figure 3). On the other hand, the Green Engineering Principles (GEP) were developed in 2003 by Anastas and Zimmerman to complement the 12 GCPs with important concepts on environmental impact, providing with a life cycle perspective [25,26]. The current scenario of environmental pollution and the need for eco-friendly actions have promoted the infusion of the Green Principles in various scientific fields, intending to limit the excessive use of resources and hazardous chemicals and abolish any potentially harmful by-products or reduce their harmfulness. Considering the economic and environmental impacts, the Green Principles are employed in governmental monetary policies, industrial and technological management, and educational systems and practices [27,28,29]. For heterogeneous photocatalysis, the main objective of the GCPs is the production of eco-friendly photocatalysts that display a wide range of light absorption and high reaction performance, using soft synthetic processes.
Heterogeneous photocatalysis to treat water and to produce green H2 has the potential to be in line with some GCPs. For example, using minimal amounts of photocatalysts to achieve the complete mineralization of organic matter in water and to continuously produce green H2 prevents waste rather than treating residues (GCP 1), which are actually produced in other process, such as adsorbents used in water treatment. Additionally, several photocatalysts are produced using soft processes, generating less hazardous chemical syntheses (GCP 3). In fact, green chemistry is embedded in the synthesis of new photocatalysts [31] to achieve safer materials (GCP 4). The development of mild-conditions and soft synthesis process are comparatively less energy-consuming and sometimes they use renewable feedstocks, such as sunlight or microbial biomass [32,33,34] (GCPs 6 and 7). Lastly, some photocatalysts are designed to be safely disposed after they are exhausted (GCP 10); bismuth-based and iron-based materials are good examples of this [35,36,37,38]. When we think about an upscaled photoreactor, the synthesis of the adequate amount of catalyst must be considered [37]. The use of safer chemicals for the large-scale synthesis of photocatalysts with harmless by-products is the primary factor to be investigated. [39]. All the GCPs are interconnected and the basic idea is to adhere to environmentally friendly and energy-efficient approaches to prevent pollution and ensure a safer implementation of the process. This may constitute the commercialization of the photocatalytic processes without harming the environment. Additionally, the GCPs contribute to the circular economy that indicates the efficient use of resources from the production to recycling, with zero waste due to 100 percent utilization of the resources. Practicing the GCPs may stabilize the circular economy, balancing the environment with economic and society development [40]. By joining the GCPs with a circular economy, it is possible to achieve sustainable development goals [41].
Green Engineering Principles (Figure 4) provide a framework for scientists and engineers to design effective, ecologically intelligent materials, products, and systems in conjunction with GCPs. So far, the published works on the application of GEPs are very scarce compared to those focused on the synthesis and reaction processes adopting the GCPs. We consider that upscaling the heterogeneous photocatalysis process should be performed under both Green Principles.
As mentioned above, heterogeneous photocatalysis prevents wastes instead of treating them (GEP 2). In terms of the degradation of organic pollutants in water, photocatalysis looks for the complete mineralization of the organic matter, transforming complex molecules into CO2, H2O, and other mineral compounds. By achieving this, more reactants are converted to the desired products (minerals) and less in undesired products (potentially toxic intermediates), addressing GEP 2. New photocatalysts are created for using less material to achieve the maximum conversion, such as ultrathin films [42,43], quantum dots, and highly efficient powder catalysts [44]. Additionally, one of the main goals of heterogeneous photocatalysis is to produce sunlight active materials, thus maximizing the mass, energy, space, and time efficiency of the whole process (GEP 4). To design photocatalytic reactors, the maximum utilization of light sources is crucial. Most photocatalytic reactors use electric light sources [45]; in this concern, electric lamps can be replaced by renewable energy harvesting modules, such as solar panels. Heterogeneous photocatalysis has evolved for the last 20 years to minimize energy consumption in separation and purification operations (GEP 3) by the implementation of (a) immobilized photocatalysts (including supported materials and thin films), (b) magnetic materials, and (c) photocatalytic membranes, among other strategies. When the lifetime of photocatalysts is over, some of them can be potential pollutants, flouting GCPs 1 and 4, as well as GEPs 1, 2, and 11. Hence, new photocatalysts should designed with an afterlife perspective, using non-hazardous materials to encourage their reuse when they are exhausted or designing materials able to be reactivated after several reaction cycles. The latter is especially useful for green H2 production as cost-effectiveness of the material increases with the number of cycles they can be used without significant performance decreases. Although some metals are abundant on the Earth’s crust, others are considered endangered and critical, as shown in Figure 5.
The availability and recyclability of raw materials must be considered when designing new photocatalysts for any purposes. For example, Pt is considered as a critical raw material for the European Commission, hence its utilization is restricted despite its outstanding properties as co-catalyst for green H2 production [46]. The photocatalysts must be designed and synthesized for being recovered after their utilization. Different processes such as solvent extraction, precipitation, ion exchange, crystallization, electrowinning, and electrodeposition are plausible methods for recovery. For example, electrodeposition is a low-cost, scalable, and straightforward method to recover bismuth-based semiconducting materials [47].
Developing new photocatalysts or upscaling the photocatalytic process must be performed following a life cycle perspective, which includes the analysis of the environmental risks carried by the synthesis, the usage, and the disposal of photocatalysts, the effluents produced after water treatment, the emissions from cradle to grave, and the overall upscaling. Life Cycle Analysis also covers the economic sustainability of the process, considering the use of electric energy and the embedded energy costs of consumable chemicals. We consider that integrating Green Principles to the synthesis of new photocatalysts will make room for a new generation of photocatalysts where Life Cycle Analysis will provide a wider picture of the photocatalytic water treatment and green H2 production systems, causing the process to be more attractive for commercialization. So far, very few works have reported the implementation of Green Principles and Life Cycle Analysis in the development of the green photocatalysis process to shed light on the current bottlenecks to implement these approaches. Therefore, further discussion must be focused on the attempts to upscale heterogeneous photocatalysis from a perspective that includes the Green Principles and Life Cycle Analysis.

3. Upscaling Heterogeneous Photocatalysis for Water Purification and Green H2 Production

Significant efforts have been undertaken for upscaling the heterogeneous photocatalysis process. The first requirement to achieve this goal is the availability of highly efficient, cost-effective, and stable photocatalysts. In this regard, a plethora of photocatalysts have been tested for the degradation of organic pollutants in water and for green H2 production since the late 1990s. Metal oxides, such as TiO2, CeO2, WO3, and bismuth-based oxides, are the most used materials (Table 2). TiO2 is the most popular material in heterogeneous photocatalysis, as it is cheap, innocuous, stable, and highly efficient. However, it is photo-excited under UV light illumination (λ < 380 nm), which limits its use in solar plants. Different strategies have been implemented to increase the photocatalytic performance of TiO2 under visible light irradiation, including doping with non-metal atoms and coupling with low bandgap semiconductors [48]. Doping semiconductors results in the formation of inter-band states, reducing the bandgap energy and thus the activation energy. However, it also increases the recombination rate of the charge carriers, which can be managed by the decoration of the semiconductor surface with nanoparticles of noble metals (e.g., Au, Pt, or Pd) [49]. For semiconductor—semiconductor junctions, type-II heterostructures are sought to boost the separation of the charge carriers. In the case of TiO2 type-II heterostructures with low-bandgap semiconductors, such as WO3, CdS, CeO2, bismuth-based, and iron-based oxides are the most reported [50]. Through the aforementioned modifications, different TiO2-based materials are currently used in sunlight-driven photocatalysis systems, displaying outstanding results [51]. In the case of WO3, CeO2, and bismuth-based semiconductors, these materials have lower bandgap values hence they are photoactive under visible light irradiation. However, the recombination of the charge carriers is considerably higher than that of TiO2, making necessary the modification of these materials. The insertion of inter-band orbitals by doping with non-metal and transition metal atoms is widely used to achieve this. Additionally, the synthesis of secondary and ternary heterostructures to transport the charge carriers from one semiconductor to another is a common way to foster the separation of charge carriers, resulting in highly efficient S- and Z-schemes. Such strategies are detailed reviewed elsewhere [52,53,54]. Different properties of the metallic oxides should be considered when they are used as photocatalysts. For example, the particle size of WO3 is normally obtained in ranges near its Bohr radius, which results in quantum confinement effects, such as a blue shift in the bandgap [55]. For tungsten- and bismuth-based materials, photocatalysis under UV light irradiation should be avoided, as these are highly reducible ions that are converted to their metallic state after few reaction cycles.
Photocatalytic reactors have been developed and tested for many years, using several configurations and materials. Different optimization strategies have been used to increase the efficiency of these systems at higher scales. Reactor configuration/geometry is a crucial factor to be considered when upscaling at microplant, pilot plant, or industrial levels [120,121,122]. Additionally, optimizing the operational parameters of the reactor can be performed using the desirability function method, which helps to achieve the efficient target response proportional to the efficiency of the upscaled system [123].
In laboratory and higher scale reaction systems, the photocatalyst is present either suspended, supported on inert particles, or as thin films; for the latter, the photocatalyst can be supported on glass, quartz, ceramics, or metallic substrates [124]. Glass is regularly used as substrate for thin films, while metallic substrates, such as stainless steel, are preferred for full-scale processes as they can cope with high mechanical stress and are inert to the reactive species produced upon the irradiation of the semiconductor [124]. On the other hand, plastic materials, such as polythene, poly acrylate, or PVC tubes are also used as substrates of thin films as they are flexible and cheap [123,125]. However, polymers can be etched by the reactive oxygen species, releasing the catalyst to the aqueous phase. This drawback can be addressed by coating the polymeric substrate with an inert layer, such as SiO2 [126]. Supporting the photocatalyst facilitates the recovery of the material to further reaction cycles, in the case of batch systems, or allows for a continuous process in H2 evolution. Removing the unitary operation to recover the catalyst leads to a more efficient use of energy, which is in line with GEP 3. However, when the photocatalyst is supported, a considerable fraction of the surface area is lost due to the agglomeration of the particles on the support, resulting in a drop of the reaction efficiency. The synthesis of photocatalytic thin films requires the deliberately production of material in excess, which is a flaw in accordance with GEP 8. Recent research has focused on the development of ultrathin films supported on different substrates, aiming at increasing the efficiency in material usage to achieve high performances in the removal of organic pollutants from water and the water splitting using less material [42,127]. Even when the immobilization of the photocatalyst can provide advantages, slurry reactors are preferred in research laboratories and mild-scale plants [128] for practical and cost-effective reasons. According to the perspective study of Toe et al. [129], a photocatalytic slurry system is nearly 12% more cost-effective compared to a panel photoreactor to produce green H2 under the case scenario of producing 10 kg H2 per day. In addition, slurry photoreactors provide better light penetration and scattering, superior mass transference, and easier installation and maintenance compared to photoreactors where the catalyst is deposited as thin films [128,129]. Conversely, using an excess of the powder photocatalyst leads to the decrease in the reaction performance because of over light scattering, causing a reduction in the quantum efficiency [130], light screening effects, and increased costs associated with the separation of the solid material. The latter is a crucial step in water treatment systems, as the remaining particles of the photocatalyst impact the quality of the effluent, especially when the semiconductor is made up by toxic elements, such as Cd, Pb, or some halides. The catalyst recovery can be performed using membranes that act as filters for the further retrieval of the solid. On the other hand, the photocatalyst can be supported on the membranes, resulting in reactive photocatalytic filters for the degradation of contaminants. In this scheme, the filtration cake formed through the process is simultaneously degraded with the dissolved contaminants [64,131]. The utilization of photocatalytic membranes and thin films synthesized by different chemical methods (e.g., dip coating, chemical vapor deposition, pulsed laser deposition, and ultrasound process) are mostly preferred over slurry photoreactors at microplant or industrial scales [128,129,130]. However, they are yet to be commercially established due to the difficulty of immobilizing the nanomaterials over large surfaces [132]. For green H2 production, the recovery of the catalyst is not necessary as the photocatalytic system is kept functioning as much as possible while the main gaseous products are recovered at the upper part of the reactor. Contrary to the definition, photocatalysts normally deteriorate as they are reused due to (a) poisoning with dissolved minerals, (b) photo-corrosion, and (c) solvation. Hence, powder and immobilized photocatalysts must be highly reusable to be considered as cost-effective materials in heterogeneous photocatalysis. According to Uekert et al. [133], a lifetime of at least one year is essential for a photocatalyst to be considered for practical application at higher scale.
So far, different reactors have been tested for upscaling purposes. The parabolic trough concentrator (PTC) was the first reactor for sunlight-driven photocatalytic water treatment [134]. Even though it has the advantages of harvesting sunlight and high mass transfer, the heating in the system is a disadvantage that may diminish the process efficiency, given that the adsorption of organic molecules on the catalyst decreases as the temperature rises. Additionally, their apertures collect only the direct incident rays, which may question the practicality of the PTCs in various climatic conditions [134]. The implementation of sun tracking systems increases the performance of the photocatalytic process, raising the prices of the whole system. Later, non-concentrating collectors (NCC), without the necessity of tracking systems, were introduced [128]. Flat plate collectors are typically used in NCCs, on which the catalyst is supported, and the liquid phase reaction mixture is directly in contact with the support before it reaches the catalyst. The NCCs are cost-effective and function in all climatic conditions as they use the sunlight beam from all directions and the tracking system is unnecessary, causing the system to be cost-effective. Compound parabolic concentrators (CPCs) are the next generation of photoreactors, combining NCC and PTC advantages. CPCs are alternatives to the reactors with the tracking system as they can collect both direct and indirect radiation due to their architecture, being an energy-saving type depending on the availability of sunlight irradiation [135]. Non-collecting reactors, such as tubular, shallow ponds, and falling film reactors have been utilized for photocatalytic water treatment [128,134,136]. Reactors using lamps as light sources have also been tested, with annular reactors as an example, and systems with lamps around the reactor chamber instead of inside [134]. So far, these systems are limited to pilot and demonstration-scale plants [128]. Overall, the upscaling process has been a challenge compared to the development of new photocatalysts (see Figure 2). This trend of being in the comfort zone of laboratory-scale synthesis and applications must be changed to boost the application of material science and thus develop the technology for upscaling photoreactors looking for the best conditions and further commercialization.

Selected Upscaling Case Studies

Upscaled photocatalytic systems have been tested mostly for water treatment worldwide. The distribution of this kind of treatment plants mainly depends on the ambient conditions, with sunlight irradiation being the most important parameter. Many of the upscaled systems are in the sunbelt region, in locations such as Tabernas in Spain, Arica in Chile, and Doha in Qatar. Some other sites, such as the Sonora and Atacama deserts, Sub-Saharan Africa, and West Australia reportedly have higher photovoltaic power potential and global horizontal sunlight irradiation (Figure 6), with a great potential for testing photocatalytic plants. Upscaled photocatalytic reactors located in high sunlight irradiation regions are greatly effective, although the economic and design parameters vary considering the inter-day and seasonal oscillations of sunlight irradiation [137].
In the National Renewable Energy Laboratory (Albuquerque, USA), six aligned PTCs were tested. The devices were modified with a UV-light transparent pipe and the reactors were irradiated with sunlight concentrated 50 times. The light reached the 450 m2 parabolic collector using a movable collector aperture perpendicular to the incident sunlight and the reflected rays were concentrated to an absorber tube aligned to the parabolic concentrator. Using this device, it was demonstrated that 1,2,3,4-tetrachloro-p-dibenzodioxin can be photocatalytically degraded using the full solar spectrum, compared to that observed when the 450 nm cutoff filter was adapted. This was the first attempt toward an upscaled photocatalytic system and the first known outdoor plant [138]. Other PTCs were reportedly installed at the Lawrence Livermore National Laboratory (Livermore, CA, USA) to treat groundwater contaminated with trichloroethylene, using a TiO2-based solar plant [139]. In this 158 m2 parabolic concentrator, 570 L of groundwater were irradiated with an effective sunlight concentration ratio of approximately 20, achieving a degradation yield of 90% in single pass experiments at neutral pH. The reaction efficiency increased to 100% under acidic conditions (pH = 5.6). The optimal concentration of TiO2 in the slurry reactor was 1 mg L−1, while increasing the flow rate marginally improved the degradation of trichloroethylene.
In the Plataforma Solar de Almería, Spain, some PTCs were built to detoxify water. The consortium of six parabolic modules was designed to treat up to 419 L of water, at flow rates from 250 to 3500 L h−1, which has been the first engineering-scale solar photochemical facility in Europe for such a purpose [140]. Most recently, a Non-Concentrating Collector (NCC), consisting of a rectangular staircase vessel with 21 steps, covered by a Pyrex glass screen to avoid evaporation, was also tested to remove 4-chlorophenol from water (Figure 7). Commercially available TiO2 was coated on non-woven paper using SiO2 as a binder agent and the surface-modified sheets were deposited on the steps of the photoreactor [141]. Different experiments were performed to treat a maximum volume of 35 L of wastewater effluent. The complete degradation of the target molecules was achieved upon 5 h of sunlight irradiation at a flow rate in the range from 1 to 7.5 L min−1, although the mineralization yield was below 100%. Still, the conversion achieved using the non-concentrating collector was slightly higher than that attained with a compound parabolic collector. Moreover, the detoxification of water containing the Congo red and indigo carmine dyes was reported in terms of the degradation of the molecules and their mineralization. This was one of the first reaction schemes demonstrating that supported photocatalysts can efficiently remove organic pollutants in water, avoiding the unitary operation of separating the solid from the liquid phase.
In recent years, different efforts to harness the abundant sunlight irradiation in Spain have been performed. For example, a flow-through reactor using stainless steel wires to support nanosized ZnO was tested at the Instituto Nacional del Carbón to remove methylene blue (10 mg L−1) from water using UV irradiation [142]. In laboratory scale experiments, a flow of 20 mL h−1 were passed through the photocatalytic mesh as the UV light source (300–400 nm, 10 mW cm−2) was 26.6 cm above. From 50 to 60% of the methylene blue was degraded after 3 h of irradiation, depending on the amount of ZnO deposited. On the other hand, the mineralization was modest, from 20 to 30% after 17 h of irradiation. Some release of ZnO from the stainless-steel mesh was observed, necessitating an increase in the adherence of the semiconductor to the support to avoid posing risks on the aquatic organisms receiving the effluents when this system is scaled.
The SOWARLA demonstration plant at the German Aerospace Center is planned to be capable of purifying water polluted with rocket fuels (some of them containing cyanide, nitrite, and hydrazine derivatives) using a new configured solar reactor, with a cost five times lower than a conventional UV-light harvesting reactor (Figure 8) [143,144,145].
A prominent example of an industrial scale photocatalysis plant to treat water using compound parabolic concentrator is SOLARDETOX, which was developed by the European Industrial Solar Consortium. With a parabolic mirror and a series of Pyrex tubes, the plant collects irradiation over 100–150 m2 to treat large volumes of water using TiO2 to remove dichloroacetic acid and cyanide [139,146]. Unfortunately, this technology has not been used to treat real wastewater streams yet.
In the United States of America, a pilot-scale photocatalytic membrane reactor (PMR) was used to remove 32 endocrine disrupting compounds at ng L−1 levels in water from the Colorado River. This patented configuration uses UV/TiO2 photocatalysis, followed by filtration through a ceramic microfiltration membrane to recover and recycle the photocatalyst [147]. In this reactor, 50 mg L−1 of the photocatalyst was exposed to UV light (λ = 185 and 254 nm) from a set of 32 lamps through very short timespans (below 30 s), at a flowrate of 24 L min−1. Steroid hormones were completely removed from water, along with most of the tested pharmaceuticals, including diclofenac, sulfamethoxazole, and fluoxetine, while flame retardants, especially perfluorooctanesulfonic acid, remained in the water.
The PMR technology has also been used in the Planta Solar de Almería, which was adapted to a compound parabolic concentrator using a TiO2 P25 suspension (0.2 g L−1) to remove lincomycin (from 10 to 50 µmol L−1) in 40 L of water at a flow rate of 10 mL s−1 and displaying promising results. The system can be used either as a batch or continuous reactor, finding the complete degradation and high mineralization of the antibiotic regardless of its initial concentration [148]. The use of membranes in the system significantly reduced the presence of degradation byproducts and photocatalyst in the effluent, resulting in a safer process.
In Japan, a Rotating Advanced Oxidation contactor (RAOC) using TiO2 supported on zeolite particles was used to remove sulfonamide pharmaceuticals from water (Figure 9). A zeolite/TiO2 composite sheet was attached to a rotating wheel and researchers were able to remove up to 96% of sulfamonomethoxine after 180 min under UV-light irradiation [149,150].
A reactor functioning with solar batteries and a Bi2O3/TiO2 thin film as photocatalyst was used to purify 40 L of wastewater from aquaculture effluents in China. The photocatalyst was able to degrade organic nitrogen compounds from water using sunlight irradiation (from 55 to 70 mW cm−2) and, during the night, UV light lamps (λ = 254 nm) were activated using batteries. After 24 h of water recirculation, the organic nitrogen compounds were removed, while the content of NH4+ and NO3 decreased by 30% and 50%, respectively [151]. In the same line, a pilot-scale Thin-Film Fixed-Bed Reactor (TFFBR) was tested to inactivate the aquaculture pathogen Aeromonas hydrophila under sunlight irradiation, using a sloping flat plate coated with TiO2 as the photocatalyst. In this homemade system, 200 mL of water was in contact with a thin film of the photocatalyst (20.5 g m−2) and exposed to sunlight for 2.5 min. At a flowrate of 4.8 L h−1, the inactivation of 1.3 log units of bacterial cells was achieved in a single pass of spring water through the reactive bed. The bacterial cells displayed injuries due to the attack of the reactive oxygen species [152].
In the African continent, the effluent from a textile factory in Menzel Temime (Tunisia) was treated in a solar pilot plant with two Flow Film Reactors (FFR), with an illuminated surface area of 50 m2 and a photonic efficiency of 15%. This reactor was designed for treating up to 1 m3 h−1 of effluent and removing recalcitrant contaminants, such as dichloroacetic acid. This scheme was chosen considering the low investment costs and energy consumption requirements [153]. According to the authors, high removal of chemical oxygen demand and total organic carbon was achieved after few reaction cycles under sunlight irradiation. The optimized scheme was planned to remove an average of 405 mg of carbon h−1 m2. Unfortunately, no reports on the functioning of this plant were found beyond 2004.
The Ramadan City in Egypt, which has many industries and obtains sunlight almost all over the year, has been recently equipped with semi-industrial constant flow photovoltaic reactors that utilize the direct sunlight irradiation to purify treated wastewater. Nanocomposites comprised by SiO2/TiO2 (2:1) supported on polymeric membranes were used as photocatalysts to degrade methyl orange and phenol (10 mg L−1, each one) as model molecules. Degradation rates of 54% and 41% were obtained for phenol and methyl orange, respectively, after 3 h under autumn sunlight irradiation (868 W m−2) [154]. According to the authors, using this system lowers the operation costs of treating 0.5 m3 d−1 of water to a third of the cost of a small photoreactor using electricity. On the other hand, in Alexandria, Egypt, where the sewage water is deposited mainly through public sewers, a homemade CPC was proposed to treat industrial effluents. In this system, TiO2 was deposited on recycled polymeric substrates (cotton, polyamide, and PET, among others) and tested to remove acid yellow 28 and methylene blue (5 mg L−1 each one) as well as chemical oxygen demand. In this scaled system, 3 L of wastewater effluent were pumped at a flow rate of 1.2 L min−1 and sunlight was irradiated (2.74 mW cm−2) all day long, starting at 9:00 AM. The complete degradation of both dyes was achieved after 8 h of irradiation using TiO2 supported on PET; however, the performance of the photocatalyst dropped with the reuse, most likely due to the affectation of the polymeric substrate [155].
Regarding the production of green H2, Schroder et al. [156] made the first attempt at a large-scale with water splitting photoreactors using natural sunlight, graphitic carbon nitride as photocatalyst, and triethanolamine as the sacrificial agent. The authors reported producing around 18 L of green H2 in one month at a hydrogen evolution reaction rate of 0.22 L kW h−1 and a maximum solar-to-hydrogen conversion of 0.12%. In another study, Maldonado et al. [157] proposed a pilot-scale solar photoreactor with 2.6% conversion efficiency using Cu/TiO2 as the photocatalyst. The system was tested using different aqueous matrices, such as water/methanol, water/glycerol, and wastewater effluents. Glycerol was the best sacrificial agent, while the poorest yields were obtained using wastewater, due to the complexity of the matrix and the hydrogen—oxygen back reaction [157]. Villa et al. [158] compared the performance of Pt/N-TiO2 and Pt/CdS-ZnS in a pilot-scale plant (25 L of capacity) for water splitting, using formic acid and organic matter in municipal wastewater as sacrificial agents. Moderate conversions were obtained using wastewater, highlighting the possibility of valorization of this waste as raw material to produce green hydrogen. After 5 h of sunlight irradiation, around 3000 µmol L−1 of H2 was produced using 0.2 g L−1 of platinized N-TiO2 and formic acid as the sacrificial agent. The yield dropped to 220 µmol L−1 when the photocatalytic process was performed using wastewater and 5 g of Pt/(CdS-ZnS) as the catalyst. Overall, the conversion efficiencies were 2.5% for Pt/N-TiO2 and 1.6% using the Pt/(CdS-ZnS) material. Almomani et al. [159] tested a solar pilot plant using carbon-doped TiO2 nanotubes as the photocatalyst under sunlight irradiation. Using 50 mg of the photocatalyst, the average rate of hydrogen production was 38.66 mmol h−1 g−1 through 50 h of sunlight irradiation, which is 1.5 times higher than the maximum rate reported for other TiO2-based photocatalysts at a lower scale so far. Ren et al. [160] tested a CPC-based photocatalytic reactor for water splitting, with a volume of 11.4 L, having a maximum hydrogen production of 1.88 L h−1 under direct solar irradiation.
In the Plataforma Solar de Almería, Arzate-Salgado et al. [161] used the Au/TiO2 photocatalyst to simultaneously produce hydrogen and degrade organic pollutants in aqueous solutions under sunlight irradiation at bench scale (27 L of capacity at a flow rate of 20 L min−1). Using a catalyst loading of 0.2 g L−1, the authors obtained 6.5 mmol H2 L−1 after 5 h of irradiation and a STH conversion efficiency of 1.8%. Even when formic acid demonstrated to be the most efficient sacrificial agent, dissolved organic matter displayed a positive effect on the photocatalytic water splitting process.
In 2018, the State Key Laboratory of Multiphase Flow in Power Engineering from the Xi’an Jiaotong University built a pilot-scale system based on CPC reactors (3.6 m2 in area and 23 L in volume). A sodium sulfate solution was used as the liquid matrix and Cd(1−x)ZnxS was the catalyst in the slurry reactor, at a loading of 2.77 g L−1 and a linear velocity of 1.2 m s−1. The average hydrogen production of 184.3 mL min−1 was achieved when the average solar radiation was 803.8 W m−2 [162].
Most recently, in 2021, Nishiyama et al. [163] built a 100 m2-scale plant for the photocatalytic water splitting. It consisted of 1600 reactor units and a gas separation facility, using SrTiO3:Al as the photocatalyst. The active volume of the system was 2.4 L. According to the authors, the system reached a STH of 0.76% and was operated during a year under field conditions with no incidents. Still, the STH conversion rate should be from 5 to 10% to ensure the systems are economically viable, meaning much work to improve the efficiency remains ahead.

4. The Foreseen Challenges in Upscaling the Heterogeneous Photocatalysis Process for Green H2 Production

Most of the reported research covers the topics related to heterogeneous photocatalysis for water treatment, while less reports are on the upscaled green H2 production via the water splitting reaction. In our opinion, this is the result of (a) the lack of effective and durable photocatalysts for the water splitting reaction; (b) the complex experimental setups that are hardly scalable; (c) the safety measures necessary to work with H2; (d) the challenges related to H2 storage; and (e) the difficulties of handling and distributing H2 from plants to the final consumer. As mentioned above, the photocatalysts used for water splitting must have a reduction potential in the conduction band higher than that necessary for the reduction in H+ into H2. A limited number of metallic oxides meet this condition and, therefore, surface modified oxides are used. For example, TiO2 and WO3 are modified with noble metal nanoparticles to reduce the recombination of charge carriers and slightly increase the reduction potential of the photo-electrons. This leads to highly effective photocatalysts [55,164]; however, the decoration with noble metal nanoparticles increases the costs of the materials and promotes the use of hazardous reagents, such as ethylenediamine. Bismuth-based semiconductors are promising materials for H2 production due to their highly reductive conduction band and the low oxidation potential of the valence band [165,166]; the latter opens the possibility of high conversion with no addition of sacrificial agent. WO3 and bismuth-based semiconductors are used in heterostructures with TiO2 for H2 production under visible light irradiation [167]. However, both oxides are readily reducible hence lowering their durability as photocatalysts under reductive conditions. When other calcogenides, such as sulfides or selenides are used, photocorrosion significantly reduces the durability of the materials, as occurs with CdS and Bi2S3. The synthesis of semiconductor—semiconductor heterostructures, removing oxygen from the photocatalytic system and decoration with noble metal nanoparticles are strategies to increase the lifetime of these photocatalysts by inhibiting photocorrosion [168]. Multiple efforts must be performed to obtain durable photocatalysts for green H2 production using synthesis processes that meet the Green Principles. The typical laboratory scale setup for the photocatalytic production of green H2 consists of a tightly sealed glass reactor, the light source, a continuous stirring system, and the gas purge. Batch configuration is mainly used, where the reagents and the catalyst are introduced into the reactor after the air is purged. The efficiencies obtained at this scale do not necessarily apply to the upscaled system. Incorporating the optimized conditions obtained at lab-scale to the upscaled system is of high importance. Homogeneity of the reaction medium is an essential factor in the upscaled systems either using slurry rectors or immobilized catalysts, which is possible through an efficient stirring system. Palette mixing systems for low-density materials and flotation to maintain a homogeneous loading of denser solids are options in this concern. In scaled systems, the light source must cover the whole surface of the reactor and guarantee the maximum number of photons reaching the bottom of the reactor to avoid poorly illuminated spots. To accomplish this, high-power lamps can be used, with the consequent increase in the reaction temperature; in such a case, a circulating cooling system is necessary to dissipate the heat. Even though immobilization is an efficient solution to reduce the complexity of the process, the fouling of the immobilized catalysts may be an issue that further affects the efficiency of the system [169].
Boiling point, ignition temperature and energy, density and permeability for different materials, high diffusion coefficient in air, and latent heat of combustion cause hydrogen to be an excellent fuel. To use this product to supply clean energy, it is necessary to consider the risks of leakage, ignition, and explosion. Hydrogen embrittlement consists of the exposition of metallic and plastic liners to H2, which may change their mechanical properties, resulting in cracks and blisters on pressurized vessels and pipelines. Leakage is hardly detected since hydrogen is colorless, odorless, and tasteless [170]. The leakage of gaseous H2 in closed spaces can cause asphyxia even at low concentrations, as it replaces the oxygen, while in both open and close spaces it may cause explosions. An initial explosion or jet fire can trigger a domino effect with nearby vessels by thermal radiation, resulting in subsequent fireballs or another jet fire. In the case of liquid H2, the main problem is the change of pressure necessary to liquefy the gas; the difference in temperature of liquid H2 and the outside evaporates the released fuel in complete or partial form. When total vaporization of the fuel occurs, it mixes with air creating a flammable cloud. In contrast, in partial vaporization, a cryogenic pool is formed on the ground surface, absorbing heat from both the atmosphere and the ground to create a boiling film while the freezing of the solid occurs followed by the condensation or freezing of the surrounding air [171]. To avoid hydrogen embrittlement, extensive research is going on over the materials of the pipes and tanks for synthesis and storage purposes to improve the reactors and devices.
Leakage of liquid H2 can cause a Boiling Liquid Expanding Vapor Explosion (BLEVE) as the leak changes the inner pressure of the H2 vessel, thus generating an expansion of the liquid and vapor phases. This causes the rupture of the vessel at atmospheric pressure and under temperature above the boiling point of the gas, which immediately turns into an expansive wave of vapor. An example of a BLEVE event and its potential consequences occurred in the San Juan Ixhuatepec explosions in Mexico in 1984. The gas plant was destroyed and over 500 persons were killed, 7000 had injuries, and nearly 60,000 people were evacuated; 149 houses were completely destroyed [172].
In terms of H2 storage, chemical and physical sorption have been studied as alternatives to store this green fuel. Some materials, such as MOFs and carbon porous materials have been studied with promising results. On the other hand, the formation of metallic hydrides allows for storage at low temperatures and pressure. The most widely tested metal hydrides are sodium alanate, lithium imide, and lithium and sodium borohydride [173]. Some of the challenges related to chemical sorption are costs, difficulties handling the storing material, and the occurrence of unwanted gases during desorption. A significant challenge is how H2 production and storage can meet the GCPs and GEPs (especially GCP 5 and 12 and GEP 1 and 2). The risks of leakage, ignition, and explosion in storage cylinders can be reduced through the implementation of different measures, such as (a) improving the design with materials resistant to embrittlement, and (b) using safety devices, such as Thermal activated Pressure Relief Devices (TPRD) [174].

5. Automation and Machine Learning as Powerful Tools to Upscale Heterogeneous Photocatalysis following the Green Principles

Automatization refers to the creation and application of a wide range of technological approaches to monitor and control the production and delivery of products with lower human intervention, increasing the precision of the process and also the massive production [175]. In the field of heterogeneous photocatalysis, automation is mostly used at the laboratory scale to (a) accelerate the synthesis and characterization of new photocatalysts and (b) the in situ or in operando monitoring of the photocatalytic reaction [176,177]. So far, automation is reported in the literature for the control of valves in green H2 production devices [178,179], the implementation of Cartesian robots in high-throughput synthesis, characterization and reaction schemes [180,181], and the generation and implementation of flexible automation tools working on lab benches to synthesize photocatalysts for green H2 production under optimal conditions [182] (Figure 10). Until 2020, the use of industrial robots in materials research laboratories was limited because of the high costs, safety issues, and the necessity of highly specialized personnel for control. However, this situation has been reverted with the drop of the costs of highly specialized robots [183].
Through the automation of the synthesis and reaction processes, data sets are rapidly generated (data mining) and analyzed to achieve the best possible synthesis and reaction conditions and further upscale the photocatalytic systems. Automation is seen as one of the bricks to build scaled photocatalysis plants and it is currently used in the first large-scale photocatalytic solar H2 production plant for the control of valves and pumps as well as to monitor the water level in the reactor [163]. On the other hand, high-throughput methods in conjunction with machine learning boost the discovery of highly active photocatalysts for solar energy application [184]. Through these approaches, some of the Green Principles can be addressed in the synthesis and characterization of new or improved photocatalysts. In this section, the applications of automation, high-through methods, and machine learning applied to heterogeneous photocatalysis are revised on the basis of their contributions to upscale the process by complying to as many Green Principles as possible.

5.1. High-Throughput Methods for the Discovery of New Photocatalysts at Laboratory Scale

High-throughput or combinatorial methods have been widely used in chemistry since the 20th century to create new bioactive molecules from a wide range of precursors [185]. In the late 1990s, high-throughput methods started to be used to discover new photocatalysts by obtaining libraries of semiconductors. These libraries were constituted of different compositions of semiconductors, including heterostructures, for screening studies aimed at finding the optimal compositions to be used in heterogeneous photocatalysis [186]. Sol-gel and solvothermal routes have been the most commonly used synthesis methods to obtain powders in high-throughput schemes, due to their simplicity, low cost, and fine control of the morphology and architecture of the obtained nanoparticles [187,188]. Several automated devices are currently available to parallelly synthesize and characterize powder semiconductors at the laboratory scale [189]. On the other hand, research on the synthesis of thin films through combinatorial methods has been more prolifically explored in the last two decades, implementing the automation for the synthesis and characterization of photocatalysts to produce green H2. For example, Kumari et al. [190] used combinatorial reactive magnetron cosputtering methods to synthesize three libraries (342 materials each one) of Fe-V-O thin films. By the variation of the composition and thickness of the films, the highest efficiency for the solar water splitting was achieved with the FeVO4 phase and a Fe composition from 54 to 66 at.%, finding no effects from the film thickness. More recently, the libraries of vanadate-based semiconductors were expanded using other metals, including Cu, Ag, W, Cr, and Co [191]. Different research groups worldwide have adapted their devices to parallelly synthesize wide libraries of thin film photocatalysts, in most cases using robotized procedures to significantly reduce the time consumption of the synthesis and characterization steps while maximizing the accuracy [192]. High-throughput synthesis at the inter-laboratory level has been developed to boost collaborative—and remote—work that adds new information for data mining to rapidly discover new photocatalysts while fostering the independent corroboration of the information generated in other research groups. In this regard, Hattrick-Simpers [193] reported high reproducibility across different laboratories for the synthesis of Zn-Sn-Ti-O thin films via high-throughput methods based on combinatorial physical vapor deposition (cosputtering and pulsed laser deposition). High-throughput methods can be also applied for the synthesis of micro-photoreactors, which have proven to efficiently increase the photocatalytic performance compared to some thin films, providing advantageous fast retrieval of the catalyst from the aqueous phase [194,195,196]. Research groups worldwide have proven the efficiency of continuous flow and immobilized microfluidic-based photo-reactors to degrade organic pollutants from water [197,198]. Most of the reported studies use bare or doped TiO2 and ZnO, with different arrays [199,200,201]. On the other hand, metal-organic cages are combined with semiconductors, resulting in visible-light active micro-photoreactors for green H2 production [202]. More recently, Claes et al. [195] reported the synthesis of translucent packed-bed TiO2 based micro-photoreactors achieving an outstanding performance to degrade methylene blue as well as energy efficiency and displaying a photocatalytic space-time yield (PSTY) of 0.657 m3 day−1 m3 reactor KW−1 under visible-light irradiation; such results are higher than that reported in previous works [196]. To achieve the upscaling of the micro-photoreactors, high-throughput methods must be used to accurately synthesize thousands of reactors in parallel for treating high volumes of water.
The miniaturization of the synthesis process in high-throughput methods promotes the compliance of GCP 1, by reducing the reagent consumption and the generation of wastes. Additionally, combinatorial methods look for increasing the atom economy (GCP 2) by optimizing the proportions of the reagents to achieve the finest composition that results in the highest photocatalytic performance. One example of this is the synthesis of g-C3N4, with a bandgap energy of ≈2.7 eV, which can be reduced by promoting the occurrence of nitrogen vacancies [203]. The synthesis of N-defective g-C3N4 is normally achieved using NH3 in excess [204]. Through combinatorial methods, the optimal composition of g-C3N4 is obtained in a miniaturized scheme, reducing the production of hazardous residues, thus achieving the atom economy to obtain the most efficient photocatalytic material that can be directly used in scaled systems.
Regarding the Green Engineering Principles, using combinational methods for the synthesis of new or improved photocatalysts prevents contamination over treatment (GEP 2) and maximizes the use of reagents and energy (GEP 4), while a minimal quantity of materials is synthesized to meet the needs of the tests prior to upscaling, thus avoiding the production of excess (GEPs 5 and 8). Lastly, using micro-photoreactors for the degradation of pollutants in water streams observes the complying of GEPs 3 and 4, as the operation is designed to minimize the use of materials and energy in a purification process.

5.2. The Application of Virtual High-Throughput Methods to Migrate from the Edisonian to the Theoretical Approach

Combinatorial synthesis and characterizations are used in tandem with computational screening studies to discover new photocatalysts or allow for more in-depth investigation of existing materials to find better performances [205]. By this approach, researchers evaluate several key parameters of a library of materials using computational tools to find the best candidates to perform the photocatalytic process. So far, DFT-based high-throughput calculations are the most commonly used to propose new photocatalysts based on their physical, chemical, and optoelectrical properties [206]. The application of large-scale computational screening methods to develop nanometric materials is possible by exploiting the capabilities of supercomputers and using the existing databases of photocatalysts, therefore this kind of studies has increased in the last decade [207]. Most of the high-throughput computational screening studies are aimed at the discovery of new photocatalysts for the water splitting process [208,209,210,211] For example, Zhang et al. [208] selected 205 layered materials over 50,000 inorganic compounds from the Materials Project Database as precursors to obtain 2D monolayers for green H2 production. By considering the bandgap value and the potential of the conduction band (0 eV vs. NHE at pH = 0), the authors reported 36 semiconductors that satisfied the conditions for photocatalytic water splitting. The list included calcogenides, arsenides, and halides (e.g., GeAs2, GeTe, TiPbO, and Mo2SBr2). Moreover, 44 type-II heterojunctions were proposed based on the following criterion: (a) the valence and conduction band position, (b) the activity for water splitting at a range of pH values, and (c) the lattice parameters of the two semiconductors. Some examples of the selected heterostructures are MoS2/MoSe2, WS2/WSe2, and MoS2/WS2. More recently, Liu et al. [212] developed a first-principles high-throughput screening across more than 1000 potential bulk piezo-photocatalytic materials for green hydrogen production, proposing a set of 16 highly promising piezo-photocatalysts based on their optoelectronic properties and band alignment tunability; this group included BiTeCl, GaN, MgTe, and SeBr, among others.
Using high-throughput computational studies, the synthesis of new photocatalytic materials is dramatically reduced, which results in zero wastes (GCP 1) and the limited production of potentially hazardous materials (GCP 3). The development of methods that predict the ab initio properties of the materials allows to look for safer photocatalysts by using less-toxic solvents, increasing the energy efficiency and preventing the formation of undesirable by-products, thus addressing GCPs 4–6 and 11. By performing high-throughput computational screening studies, material researchers migrate from the Edisoninan approach, characterized by the trial-and-error discovery, toward the systematic theoretical approach. The latter is in line with some of the Green Engineering Principles, by developing products (materials), procedures (synthesis and characterization), and systems (photocatalysis tests) output pulled rather than input pushed (GEP 5). This is because only the most effective—at least theoretically—photocatalysts are synthesized in the exact amount required to produce the maximum conversion, preventing wastes (GEPs 2 and 8) and maximizing mass, energy, and time consumption (GEP 4).
Machine learning, a subgroup of artificial intelligence, extracts the data of the synthesis and characteristics of photocatalytic materials from the high-throughput schemes to feed a set of algorithms and build models that ultimately make predictions on the performance of the photocatalysts. The stages for constructing machine learning models are: (a) data collection to form a training data set; (b) create and choose mathematical descriptors that encode the characteristics of the materials; (c) select a proper algorithm to build the model; and (d) to assess the quality and predictive capacity of the model [213]. Machine learning represents a paradigm shift in how new photocatalysts are designed, synthesized, and tested at the laboratory scale [214], which inherently covers several of the Green Principles. Machine learning finds its finest point when is coupled to automation, as it can be jointly used to upscale the synthesis of photocatalysts. For example, Tao et al. [215] integrated machine learning and microfluidics to accelerate the identification and optimization of the reaction conditions to obtain nanosized photocatalysts. The authors performed 160 experiments in a microfluidics platform, using different synthesis protocols to produce Au nanoparticles by green routes, while machine learning algorithms found the relationship between the synthesis conditions and the reactivity of the formed nanoparticles. There are so many challenges to face in the way ahead, but we consider that the next generation of photocatalyst will be produced, tested, and upscaled following the Green Principles and with a great support from computational chemistry.

6. Perspectives

Heterogeneous photocatalysis is a promising process in environmental remediation and energy generation and there are many unexplored possibilities in this area of research. Through the years, many attempts have been performed to upscale the photocatalytic water treatment; nevertheless, very few works have been published showing the settlement of the research in the production of new photocatalytic materials at higher scale. One reason is the inadequate funding for applied investigation in most countries; changing this paradigm is one of the most challenging tasks for the future and can be achieved through the continuous efforts of upscaling in universities and research centers that foster the interest of governments and funding associations to provide more resources in this matter. Sunlight harvesting using heterogeneous photocatalysis can be a reality in countries where sunlight irradiation is available throughout the year; therefore, some funding programs promoted by developed countries to implement technological advances in developing countries can aim at using upscaled photocatalytic systems in impoverished zones. Some upscaled systems have been successfully tested in developed countries such as Germany, Spain, and China. It is time that this kind of technology reaches densely populated dry zones where the necessity for safe water is high; this will enable science and technology to be applied to address social problems related to water scarcity and green fuel production.
Green Principles must be integrated into the heterogeneous photocatalysis process from a cradle-to-grave perspective, considering the pollution produced from the photocatalyst synthesis, its use through several reaction cycles, and its disposal when their lifetime is over. The GEPs should be especially considered when scaled systems are proposed and used either for water treatment or green H2 production. Each year more and more publications include the term “green”; however, this term should be moderated and adjusted only for materials that meet most of the Green Principles and Life Cycle Analysis. Following the GCPs and GEPs is the way ahead, which seems challenging. New problems arrive when these principles are applied since new variables are integrated. Indeed, more research and safety measures are necessary to upscale the water-splitting process. Future investigations should take this process from the laboratory in industrial plants; the potential revenues foreseen in the green fuel industry could promote new research in this unexplored field.

Author Contributions

F.A.-R., J.C.D.-Á., K.T.D. and R.Z. contributed to the writing and reviewing process as well as to data curation. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Secretaría de Educación, Ciencia, Tecnología e Innovación de la Ciudad de México, grant number SECITI/047/2016.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Long, Z.; Li, Q.; Wei, T.; Zhang, G.; Ren, Z. Historical Development and Prospects of Photocatalysts for Pollutant Removal in Water. J. Hazard. Mater. 2020, 395, 122599. [Google Scholar] [CrossRef] [PubMed]
  2. Ibhadon, A.O.; Fitzpatrick, P. Heterogeneous Photocatalysis: Recent Advances and Applications. Catalysis 2013, 3, 189–218. [Google Scholar] [CrossRef] [Green Version]
  3. Singh, R.; Dutta, S. A Review on H2 Production through Photocatalytic Reactions Using TiO2/TiO2-Assisted Catalysts. Fuel 2018, 220, 607–620. [Google Scholar] [CrossRef]
  4. Singh, S.; Kumar, V.; Dhanjal, D.S.; Datta, S.; Kaur, S.; Romero, R.; Singh, J. Degradation of Pesticides in Wastewater Using Heterogeneous Photocatalysis. In Advanced Oxidation Processes for Effluent Treatment Plants; Shah, M.P., Ed.; Elsevier BV: Amsterdam, The Netherlands, 2021; pp. 161–175. [Google Scholar]
  5. Piccirillo, C.; Castro, P.M.L. Calcium Hydroxyapatite-Based Photocatalysts for Environment Remediation: Characteristics, Performances and Future Perspectives. J. Environ. Manag. 2017, 193, 79–91. [Google Scholar] [CrossRef] [PubMed]
  6. Hou, H.; Wang, L.; Gao, F.; Yang, X.; Yang, W. BiVO4@TiO2 Core-Shell Hybrid Mesoporous Nanofibers towards Efficient Visible-Light-Driven Photocatalytic Hydrogen Production. J. Mater. Chem. C 2019, 7, 7858–7864. [Google Scholar] [CrossRef]
  7. Bacha, A.U.R.; Nabi, I.; Fu, Z.; Li, K.; Cheng, H.; Zhang, L. A Comparative Study of Bismuth-Based Photocatalysts with Titanium Dioxide for Perfluorooctanoic Acid Degradation. Chin. Chem. Lett. 2019, 30, 2225–2230. [Google Scholar] [CrossRef]
  8. Zhou, P.; Yu, J.; Jaroniec, M. All-Solid-State Z-Scheme Photocatalytic Systems. Adv. Mater. 2014, 26, 4920–4935. [Google Scholar] [CrossRef]
  9. Dong, S.; Feng, J.; Fan, M.; Pi, Y.; Hu, L.; Han, X.; Liu, M.; Sun, J.; Sun, J. Recent Developments in Heterogeneous Photocatalytic Water Treatment Using Visible Light-Responsive Photocatalysts: A Review. RSC Adv. 2015, 5, 14610–14630. [Google Scholar] [CrossRef]
  10. Li, Y.; Tsang, S.C.E. Recent Progress and Strategies for Enhancing Photocatalytic Water Splitting. Mater. Today Sustain. 2020, 9, 100032. [Google Scholar] [CrossRef]
  11. Mamiyev, Z.; Balayeva, N.O. Metal Sulfide Photocatalysts for Hydrogen Generation: A Review of Recent Advances. Catalysts 2022, 12, 11316. [Google Scholar] [CrossRef]
  12. Mondal, K.; Sharma, A. Recent Advances in the Synthesis and Application of Photocatalytic Metal-Metal Oxide Core-Shell Nanoparticles for Environmental Remediation and Their Recycling Process. RSC Adv. 2016, 6, 83589–83612. [Google Scholar] [CrossRef]
  13. Amenta, R.; Amenta, S.C. Photocatalysis: Principles and Applications; CRC Press: New York, NY, USA, 2017. [Google Scholar]
  14. Acar, C.; Dincer, I.; Naterer, G.F. Review of Photocatalytic Water-Splitting Methods for Sustainable Hydrogen Production. Int. J. Energy Res. 2016, 40, 1449–1473. [Google Scholar] [CrossRef]
  15. He, X.; Wang, F.; Wallington, T.J.; Shen, W.; Melaina, M.W.; Kim, H.C.; De Kleine, R.; Lin, T.; Zhang, S.; Keoleian, G.A.; et al. Well-to-Wheels Emissions, Costs, and Feedstock Potentials for Light-Duty Hydrogen Fuel Cell Vehicles in China in 2017 and 2030. Renew. Sustain. Energy Rev. 2021, 137, 110477. [Google Scholar] [CrossRef]
  16. Ji, M.; Wang, J. Review and Comparison of Various Hydrogen Production Methods Based on Costs and Life Cycle Impact Assessment Indicators. Int. J. Hydrogen Energy 2021, 46, 38612–38635. [Google Scholar] [CrossRef]
  17. Wang, Q.; Han, L. Hydrogen Production. In Handbook of Climate Change Mitigation and Adaptation; Springer: Cham, Switzerland, 2022; pp. 1855–1900. [Google Scholar]
  18. Zoppi, G.; Pipitone, G.; Gruber, H.; Weber, G.; Reichhold, A.; Pirone, R.; Bensaid, S. Aqueous Phase Reforming of Pilot-Scale Fischer-Tropsch Water Effluent for Sustainable Hydrogen Production. Catal. Today 2021, 367, 239–247. [Google Scholar] [CrossRef]
  19. Zoppi, G.; Pipitone, G.; Pirone, R.; Bensaid, S. Aqueous Phase Reforming Process for the Valorization of Wastewater Streams: Application to Different Industrial Scenarios. Catal. Today 2022, 387, 224–236. [Google Scholar] [CrossRef]
  20. Tak, S.S.; Shetye, O.; Muley, O.; Jaiswal, H.; Malik, S.N. Emerging Technologies for Hydrogen Production from Wastewater. Int. J. Hydrogen Energy 2022, 47, 37282–37301. [Google Scholar] [CrossRef]
  21. Kayfeci, M.; Keçebaş, A.; Bayat, M. Hydrogen Production. In Solar Hydrogen Production: Processes, Systems and Technologies; Calise, F., Dentice D’Accadia, M., Santarelli, M., Lanzini, A., Ferrero, D., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 45–83. [Google Scholar]
  22. Frowijn, L.S.F.; van Sark, W.G.J.H.M. Analysis of Photon-Driven Solar-to-Hydrogen Production Methods in the Netherlands. Sustain. Energy Technol. Assess. 2021, 48, 101631. [Google Scholar] [CrossRef]
  23. Avcıoğlu, C.; Avcıoğlu, S.; Bekheet, M.F.; Gurlo, A. Solar Hydrogen Generation Using Niobium-Based Photocatalysts: Design Strategies, Progress, and Challenges. Mater. Today Energy 2022, 24, 100936. [Google Scholar] [CrossRef]
  24. Xie, W.; Li, T.; Tiraferri, A.; Drioli, E.; Figoli, A.; Crittenden, J.C.; Liu, B. Toward the Next Generation of Sustainable Membranes from Green Chemistry Principles. ACS Sustain. Chem. Eng. 2021, 9, 50–75. [Google Scholar] [CrossRef]
  25. Zimmerman, J.B.; Anastas, P.T.; Erythropel, H.C.; Leitner, W. Designing for a Green Chemistry Future. Science 2020, 367, 397–400. [Google Scholar] [CrossRef] [PubMed]
  26. Tang, S.Y.; Bourne, R.A.; Smith, R.L.; Poliakoff, M. The 24 Principles of Green Engineering and Green Chemistry: “IMPROVEMENTS PRODUCTIVELY”. Green Chem. 2008, 10, 268–269. [Google Scholar] [CrossRef]
  27. Listyarini, R.V.; Pamenang, F.D.N.; Harta, J.; Wijayanti, L.W.; Asy’ari, M.; Lee, W. The Integration of Green Chemistry Principles into Small Scale Chemistry Practicum for Senior High School Students. J. Pendidik. IPA Indones. 2019, 8, 371–378. [Google Scholar] [CrossRef]
  28. Chen, T.L.; Kim, H.; Pan, S.Y.; Tseng, P.C.; Lin, Y.P.; Chiang, P.C. Implementation of Green Chemistry Principles in Circular Economy System towards Sustainable Development Goals: Challenges and Perspectives. Sci. Total Environ. 2020, 716, 136998. [Google Scholar] [CrossRef] [PubMed]
  29. Grieger, K.; Leontyev, A. Promoting Student Awareness of Green Chemistry Principles via Student-Generated Presentation Videos. J. Chem. Educ. 2020, 97, 2657–2663. [Google Scholar] [CrossRef]
  30. Çelik, D.; Yıldız, M. Investigation of Hydrogen Production Methods in Accordance with Green Chemistry Principles. Int. J. Hydrogen Energy 2017, 42, 23395–23401. [Google Scholar] [CrossRef]
  31. Soltys, L.; Olkhovyy, O.; Tatarchuk, T.; Naushad, M. Green Synthesis of Metal and Metal Oxide Nanoparticles: Principles of Green Chemistry and Raw Materials. Magnetochemistry 2021, 7, 145. [Google Scholar] [CrossRef]
  32. Jayapriya, M.; Dhanasekaran, D.; Arulmozhi, M.; Nandhakumar, E.; Senthilkumar, N.; Sureshkumar, K. Green Synthesis of Silver Nanoparticles Using Piper Longum Catkin Extract Irradiated by Sunlight: Antibacterial and Catalytic Activity. Res. Chem. Intermed. 2019, 45, 3617–3631. [Google Scholar] [CrossRef]
  33. Jacob, J.M.; Lens, P.N.L.; Balakrishnan, R.M. Microbial Synthesis of Chalcogenide Semiconductor Nanoparticles: A Review. Microb. Biotechnol. 2015, 9, 11–21. [Google Scholar] [CrossRef]
  34. Zikalala, N.; Matshetshe, K.; Parani, S.; Oluwafemi, O.S. Biosynthesis Protocols for Colloidal Metal Oxide Nanoparticles. Nano-Struct. Nano-Objects 2018, 16, 288–299. [Google Scholar] [CrossRef]
  35. Gao, P.; Yang, Y.; Yin, Z.; Kang, F.; Fan, W.; Sheng, J.; Feng, L.; Liu, Y.; Du, Z.; Zhang, L. A Critical Review on Bismuth Oxyhalide Based Photocatalysis for Pharmaceutical Active Compounds Degradation: Modifications, Reactive Sites, and Challenges. J. Hazard. Mater. 2021, 412, 125186. [Google Scholar] [CrossRef]
  36. Roy, S.D.; Das, K.C.; Dhar, S.S. Conventional to Green Synthesis of Magnetic Iron Oxide Nanoparticles; Its Application as Catalyst, Photocatalyst and Toxicity: A Short Review. Inorg. Chem. Commun. 2021, 134, 109050. [Google Scholar] [CrossRef]
  37. Manjunatha, A.D.; Pavithra, N.S.; Marappa, S.; Prashanth, S.A.; Nagaraju, G.; Puttaswamy. Green Synthesis of Flower-like BiVO4 Nanoparticles by Solution Combustion Method Using Lemon (Citrus Limon) Juice as a Fuel: Photocatalytic and Electrochemical Study. ChemistrySelect 2018, 3, 13456–13463. [Google Scholar] [CrossRef] [Green Version]
  38. Garg, S.; Yadav, M.; Chandra, A.; Sapra, S.; Gahlawat, S.; Ingole, P.P.; Todea, M.; Bardos, E.; Pap, Z.; Hernadi, K. Facile Green Synthesis of BiOBr Nanostructures with Superior Visible-Light-Driven Photocatalytic Activity. Materials 2018, 11, 1273. [Google Scholar] [CrossRef] [Green Version]
  39. Prakash, M.; Kavitha, H.P.; Abinaya, S.; Vennila, J.P.; Lohita, D. Green Synthesis of Bismuth Based Nanoparticles and Its Applications—A Review. Sustain. Chem. Pharm. 2022, 25, 100547. [Google Scholar] [CrossRef]
  40. Velenturf, A.P.M.; Purnell, P. Principles for a Sustainable Circular Economy. Sustain. Prod. Consum. 2021, 27, 1437–1457. [Google Scholar] [CrossRef]
  41. Guo, Z.; Wang, A.; Wang, W.Y.; Zhao, Y.L.; Chiang, P.C. Implementing Green Chemistry Principles for Circular Economy towards Sustainable Development Goals. Chem. Eng. Trans. 2021, 88, 955–960. [Google Scholar] [CrossRef]
  42. Xu, Y.; Zhang, L.; Yin, M.; Xie, D.; Chen, J.; Yin, J.; Fu, Y.; Zhao, P.; Zhong, H.; Zhao, Y.; et al. Ultrathin G-C3N4 Films Supported on Attapulgite Nanofibers with Enhanced Photocatalytic Performance. Appl. Surf. Sci. 2018, 440, 170–176. [Google Scholar] [CrossRef]
  43. Ma, Y.; Jiang, X.; Sun, R.; Yang, J.; Jiang, X.; Liu, Z.; Xie, M.; Xie, E.; Han, W. Z-Scheme Bi2O2.33/Bi2S3 Heterojunction Nanostructures for Photocatalytic Overall Water Splitting. Chem. Eng. J. 2020, 382, 123020. [Google Scholar] [CrossRef]
  44. Tasleem, S.; Tahir, M. Current Trends in Strategies to Improve Photocatalytic Performance of Perovskites Materials for Solar to Hydrogen Production. Renew. Sustain. Energy Rev. 2020, 132, 110073. [Google Scholar] [CrossRef]
  45. Deng, F.; Shi, H.; Guo, Y.; Luo, X.; Zhou, J. Engineering Paths of Sustainable and Green Photocatalytic Degradation Technology for Pharmaceuticals and Organic Contaminants of Emerging Concern. Curr. Opin. Green Sustain. Chem. 2021, 29, 100465. [Google Scholar] [CrossRef]
  46. Liu, J.; Li, Y.; Zhou, X.; Jiang, H.; Yang, H.G.; Li, C. Positively Charged Pt-Based Cocatalysts: An Orientation for Achieving Efficient Photocatalytic Water Splitting. J. Mater. Chem. A 2020, 8, 17–26. [Google Scholar] [CrossRef]
  47. Thanu, V.R.C.; Jayakumar, M. Electrochemical Recovery of Antimony and Bismuth from Spent Electrolytes. Sep. Purif. Technol. 2020, 235, 116169. [Google Scholar] [CrossRef]
  48. Lee, S.Y.; Park, S.J. TiO2photocatalyst for Water Treatment Applications. J. Ind. Eng. Chem. 2013, 19, 1761–1769. [Google Scholar] [CrossRef]
  49. Durán-Álvarez, J.C.; Santiago, A.L.; Ramírez-Ortega, D.; Acevedo-Peña, P.; Castillón, F.; Ramírez-Zamora, R.M.; Zanella, R. Surface Modification of B–TiO2 by Deposition of Au Nanoparticles to Increase Its Photocatalytic Activity under Simulated Sunlight Irradiation. J. Sol-Gel Sci. Technol. 2018, 88, 474–487. [Google Scholar] [CrossRef]
  50. Del Angel, R.; Durán-Álvarez, J.C.; Zanella, R. TiO2-Low Band Gap Semiconductor Heterostructures for Water Treatment Using Sunlight-Driven Photocatalysis. In Titanium Dioxide: Material for a Sustainable Environment; Yang, D., Ed.; InTechOpen: London, UK, 2018; pp. 305–330. ISBN 978-1-78923-327-8. [Google Scholar]
  51. Lettieri, S.; Pavone, M.; Fioravanti, A.; Amato, L.S.; Maddalena, P. Charge Carrier Processes and Optical Properties in TiO2 and TiO2-Based Heterojunction Photocatalysts: A Review. Materials 2021, 14, 1645. [Google Scholar] [CrossRef]
  52. Iqbal, J.; Shah, N.S.; Khan, Z.U.H.; Rizwan, M.; Murtaza, B.; Jamil, F.; Shah, A.; Ullah, A.; Nazzal, Y.; Howari, F. Visible Light Driven Doped CeO2 for the Treatment of Pharmaceuticals in Wastewater: A Review. J. Water Process Eng. 2022, 49, 103130. [Google Scholar] [CrossRef]
  53. Kalanoor, B.S.; Seo, H.; Kalanur, S.S. Multiple Ion Doping in BiVO4 as an Effective Strategy of Enhancing Photoelectrochemical Water Splitting: A Review. Mater. Sci. Energy Technol. 2021, 4, 317–328. [Google Scholar] [CrossRef]
  54. Liao, M.; Su, L.; Deng, Y.; Xiong, S.; Tang, R.; Wu, Z.; Ding, C.; Yang, L.; Gong, D. Strategies to Improve WO3-Based Photocatalysts for Wastewater Treatment: A Review. J. Mater. Sci. 2021, 56, 14416–14447. [Google Scholar] [CrossRef]
  55. Durán-Álvarez, J.C.; Del Angel, R.; Ramírez-Ortega, D.; Guerrero-Araque, D.; Zanella, R. An Alternative Method for the Synthesis of Functional Au/WO3 Materials and Their Use in the Photocatalytic Production of Hydrogen. Catal. Today 2018, 341, 49–58. [Google Scholar] [CrossRef]
  56. Faisal, M.; Abu Tariq, M.; Muneer, M. Photocatalysed Degradation of Two Selected Dyes in UV-Irradiated Aqueous Suspensions of Titania. Dyes Pigments 2007, 72, 233–239. [Google Scholar] [CrossRef]
  57. Aljuboury, D.A.D.A.; Palaniandy, P.; Aziz, H.B.A.; Feroz, S.; Amr, S.S.A. Evaluating Photo-Degradation of COD and TOC in Petroleum Refinery Wastewater by Using TiO2/ZnO Photo-Catalyst. Water Sci. Technol. 2016, 74, 1312–1325. [Google Scholar] [CrossRef]
  58. Nassoko, D.; Li, Y.F.; Wang, H.; Li, J.L.; Li, Y.Z.; Yu, Y. Nitrogen-Doped TiO2 Nanoparticles by Using EDTA as Nitrogen Source and Soft Template: Simple Preparation, Mesoporous Structure, and Photocatalytic Activity under Visible Light. J. Alloys Compd. 2012, 540, 228–235. [Google Scholar] [CrossRef]
  59. Faramarzpour, M.; Vossoughi, M.; Borghei, M. Photocatalytic Degradation of Furfural by Titania Nanoparticles in a Floating-Bed Photoreactor. Chem. Eng. J. 2009, 146, 79–85. [Google Scholar] [CrossRef]
  60. Sboui, M.; Nsib, M.F.; Rayes, A.; Swaminathan, M.; Houas, A. TiO2–PANI/Cork Composite: A New Floating Photocatalyst for the Treatment of Organic Pollutants under Sunlight Irradiation. J. Environ. Sci. 2017, 60, 3–13. [Google Scholar] [CrossRef]
  61. Magalhães, F.; Lago, R.M. Floating Photocatalysts Based on TiO2 Grafted on Expanded Polystyrene Beads for the Solar Degradation of Dyes. Sol. Energy 2009, 83, 1521–1526. [Google Scholar] [CrossRef]
  62. Magalhães, F.; Moura, F.C.C.; Lago, R.M. TiO2/LDPE Composites: A New Floating Photocatalyst for Solar Degradation of Organic Contaminants. Desalination 2011, 276, 266–271. [Google Scholar] [CrossRef]
  63. Shavisi, Y.; Sharifnia, S.; Zendehzaban, M.; Mirghavami, M.L.; Kakehazar, S. Application of Solar Light for Degradation of Ammonia in Petrochemical Wastewater by a Floating TiO2/LECA Photocatalyst. J. Ind. Eng. Chem. 2014, 20, 2806–2813. [Google Scholar] [CrossRef]
  64. Alias, S.S.; Harun, Z.; Azhar, F.H.; Ibrahim, S.A.; Johar, B. Comparison between Commercial and Synthesised Nano Flower-like Rutile TiO2 Immobilised on Green Super Adsorbent towards Dye Wastewater Treatment. J. Clean. Prod. 2020, 251, 119448. [Google Scholar] [CrossRef]
  65. Mohammadi, M.; Sabbaghi, S. Photo-Catalytic Degradation of 2,4-DCP Wastewater Using MWCNT/TiO2 Nano-Composite Activated by UV and Solar Light. Environ. Nanotechnol. Monit. Manag. 2014, 1–2, 24–29. [Google Scholar] [CrossRef]
  66. Natarajan, T.S.; Natarajan, K.; Bajaj, H.C.; Tayade, R.J. Enhanced Photocatalytic Activity of Bismuth-Doped TiO2 Nanotubes under Direct Sunlight Irradiation for Degradation of Rhodamine B Dye. J. Nanoparticle Res. 2013, 15, 1669. [Google Scholar] [CrossRef]
  67. Peng, Y.P.; Yeh, Y.T.; Shah, S.I.; Huang, C.P. Concurrent Photoelectrochemical Reduction of CO2 and Oxidation of Methyl Orange Using Nitrogen-Doped TiO2. Appl. Catal. B Environ. 2012, 123–124, 414–423. [Google Scholar] [CrossRef]
  68. Wang, X.; Wang, W.; Wang, X.; Zhao, J.; Zhang, J.; Song, J. Insight into Visible Light-Driven Photocatalytic Degradation of Diesel Oil by Doped TiO2-PS Floating Composites. Environ. Sci. Pollut. Res. 2016, 23, 18145–18153. [Google Scholar] [CrossRef] [PubMed]
  69. Song, J.; Wang, X.; Bu, Y.; Zhang, J.; Wang, X.; Huang, J.; Chen, J.; Zhao, J. Preparation, Characterization, and Photocatalytic Activity Evaluation of Fe–N-Codoped TiO2/Fly Ash Cenospheres Floating Photocatalyst. Environ. Sci. Pollut. Res. 2016, 23, 22793–22802. [Google Scholar] [CrossRef] [PubMed]
  70. Allard, M.M.; Merlos, S.N.; Springer, B.N.; Cooper, J.; Zhang, G.; Boskovic, D.S.; Kwon, S.R.; Nick, K.E.; Perry, C.C. Role of TiO2 Anatase Surface Morphology on Organophosphorus Interfacial Chemistry. J. Phys. Chem. C 2018, 122, 29237–29248. [Google Scholar] [CrossRef]
  71. Sivagami, K.; Vikraman, B.; Krishna, R.R.; Swaminathan, T. Chlorpyrifos and Endosulfan Degradation Studies in an Annular Slurry Photo Reactor. Ecotoxicol. Environ. Saf. 2016, 134, 327–331. [Google Scholar] [CrossRef]
  72. Konstantinou, I.K.; Sakellarides, T.M.; Sakkas, V.A.; Albanis, T.A. Photocatalytic Degradation of Selected S-Triazine Herbicides and Oganophosphorus Insecticides over Aqueous TiO2 Suspensions. Environ. Sci. Technol. 2001, 35, 398–405. [Google Scholar] [CrossRef]
  73. Vaiano, V.; Sacco, O.; Sannino, D. Electric Energy Saving in Photocatalytic Removal of Crystal Violet Dye through the Simultaneous Use of Long-Persistent Blue Phosphors, Nitrogen-Doped TiO2 and UV-Light Emitting Diodes. J. Clean. Prod. 2019, 210, 1015–1021. [Google Scholar] [CrossRef]
  74. Matias, M.L.; Morais, M.; Pimentel, A.; Vasconcelos, F.X.; Reis Machado, A.S.; Rodrigues, J.; Fortunato, E.; Martins, R.; Nunes, D. Floating TiO2-Cork Nano-Photocatalysts for Water Purification Using Sunlight. Sustainability 2022, 14, 9645. [Google Scholar] [CrossRef]
  75. Djellabi, R.; Yang, B.; Adeel Sharif, H.M.; Zhang, J.; Ali, J.; Zhao, X. Sustainable and Easy Recoverable Magnetic TiO2-Lignocellulosic Biomass@Fe3O4 for Solar Photocatalytic Water Remediation. J. Clean. Prod. 2019, 233, 841–847. [Google Scholar] [CrossRef]
  76. Li, H.; Liu, J.; Qian, J.; Li, Q.; Yang, J. Preparation of Bi-Doped TiO2 Nanoparticles and Their Visible Light Photocatalytic Performance. Cuihua Xuebao/Chin. J. Catal. 2014, 35, 1578–1589. [Google Scholar] [CrossRef]
  77. Lu, D.; Yang, M.; Fang, P.; Li, C.; Jiang, L. Enhanced Photocatalytic Degradation of Aqueous Phenol and Cr(VI) over Visible-Light-Driven TbxOy Loaded TiO2-Oriented Nanosheets. Appl. Surf. Sci. 2017, 399, 167–184. [Google Scholar] [CrossRef]
  78. Barakat, M.A.; Anjum, M.; Kumar, R.; Alafif, Z.O.; Oves, M.; Ansari, M.O. Design of Ternary Ni(OH)2/Graphene Oxide/TiO2 Nanocomposite for Enhanced Photocatalytic Degradation of Organic, Microbial Contaminants, and Aerobic Digestion of Dairy Wastewater. J. Clean. Prod. 2020, 258, 120588. [Google Scholar] [CrossRef]
  79. Lai, C.; Wang, M.M.; Zeng, G.M.; Liu, Y.G.; Huang, D.L.; Zhang, C.; Wang, R.Z.; Xu, P.; Cheng, M.; Huang, C.; et al. Synthesis of Surface Molecular Imprinted TiO2/Graphene Photocatalyst and Its Highly Efficient Photocatalytic Degradation of Target Pollutant under Visible Light Irradiation. Appl. Surf. Sci. 2016, 390, 368–376. [Google Scholar] [CrossRef]
  80. Juabrum, S.; Chankhanittha, T.; Nanan, S. Hydrothermally Grown SDS-Capped ZnO Photocatalyst for Degradation of RR141 Azo Dye. Mater. Lett. 2019, 245, 1–5. [Google Scholar] [CrossRef]
  81. El Bekkali, C.; Bouyarmane, H.; El Karbane, M.; Masse, S.; Saoiabi, A.; Coradin, T.; Laghzizil, A. Zinc Oxide-Hydroxyapatite Nanocomposite Photocatalysts for the Degradation of Ciprofloxacin and Ofloxacin Antibiotics. Colloids Surf. A Physicochem. Eng. Asp. 2018, 539, 364–370. [Google Scholar] [CrossRef] [Green Version]
  82. Kaur, A.; Gupta, G.; Ibhadon, A.O.; Salunke, D.B.; Sinha, A.S.K.; Kansal, S.K. A Facile Synthesis of Silver Modified ZnO Nanoplates for Efficient Removal of Ofloxacin Drug in Aqueous Phase under Solar Irradiation. J. Environ. Chem. Eng. 2018, 6, 3621–3630. [Google Scholar] [CrossRef]
  83. Yu, Y.; Yao, B.; He, Y.; Cao, B.; Ren, Y.; Sun, Q. Piezo-Enhanced Photodegradation of Organic Pollutants on Ag3PO4/ZnO Nanowires Using Visible Light and Ultrasonic. Appl. Surf. Sci. 2020, 528, 146819. [Google Scholar] [CrossRef]
  84. Boutra, B.; Güy, N.; Özacar, M.; Trari, M. Magnetically Separable MnFe2O4/TA/ZnO Nanocomposites for Photocatalytic Degradation of Congo Red under Visible Light. J. Magn. Magn. Mater. 2020, 497, 165994. [Google Scholar] [CrossRef]
  85. Mirzaei, A.; Yerushalmi, L.; Chen, Z.; Haghighat, F. Photocatalytic Degradation of Sulfamethoxazole by Hierarchical Magnetic ZnO@g-C3N4: RSM Optimization, Kinetic Study, Reaction Pathway and Toxicity Evaluation. J. Hazard. Mater. 2018, 359, 516–526. [Google Scholar] [CrossRef]
  86. Talinungsang; Paul, N.; Dhar Purkayastha, D.; Krishna, M.G. TiO2/SnO2 and SnO2/TiO2 Heterostructures as Photocatalysts for Degradation of Stearic Acid and Methylene Blue under UV Irradiation. Superlattices Microstruct. 2019, 129, 105–114. [Google Scholar] [CrossRef]
  87. Rajendran, S.; Khan, M.M.; Gracia, F.; Qin, J.; Gupta, V.K.; Arumainathan, S. Ce3+-Ion-Induced Visible-Light Photocatalytic Degradation and Electrochemical Activity of ZnO/CeO2 Nanocomposite. Sci. Rep. 2016, 6, 31641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Sawant, S.Y.; Cho, M.H. Facile Electrochemical Assisted Synthesis of ZnO/Graphene Nanosheets with Enhanced Photocatalytic Activity. RSC Adv. 2015, 5, 97788–97797. [Google Scholar] [CrossRef]
  89. Eskizeybek, V.; Sari, F.; Gülce, H.; Gülce, A.; Avci, A. Preparation of the New Polyaniline/ZnO Nanocomposite and Its Photocatalytic Activity for Degradation of Methylene Blue and Malachite Green Dyes under UV and Natural Sun Lights Irradiations. Appl. Catal. B Environ. 2012, 119–120, 197–206. [Google Scholar] [CrossRef]
  90. Mousavi-Kamazani, M.; Rahmatolahzadeh, R.; Beshkar, F. Facile Solvothermal Synthesis of CeO2–CuO Nanocomposite Photocatalyst Using Novel Precursors with Enhanced Photocatalytic Performance in Dye Degradation. J. Inorg. Organomet. Polym. Mater. 2017, 27, 1342–1350. [Google Scholar] [CrossRef]
  91. Liu, G.; Wang, H.; Chen, D.; Dai, C.; Zhang, Z.; Feng, Y. Photodegradation Performances and Transformation Mechanism of Sulfamethoxazole with CeO2/CN Heterojunction as Photocatalyst. Sep. Purif. Technol. 2020, 237, 116329. [Google Scholar] [CrossRef]
  92. Pudukudy, M.; Jia, Q.; Yuan, J.; Megala, S.; Rajendran, R.; Shan, S. Influence of CeO2 Loading on the Structural, Textural, Optical and Photocatalytic Properties of Single-Pot Sol-Gel Derived Ultrafine CeO2/TiO2 Nanocomposites for the Efficient Degradation of Tetracycline under Visible Light Irradiation. Mater. Sci. Semicond. Process. 2020, 108, 104891. [Google Scholar] [CrossRef]
  93. Wen, X.J.; Niu, C.G.; Zhang, L.; Liang, C.; Zeng, G.M. A Novel Ag2O/CeO2 Heterojunction Photocatalysts for Photocatalytic Degradation of Enrofloxacin: Possible Degradation Pathways, Mineralization Activity and an in Depth Mechanism Insight. Appl. Catal. B Environ. 2018, 221, 701–714. [Google Scholar] [CrossRef]
  94. Murugadoss, G.; Kumar, D.D.; Kumar, M.R.; Venkatesh, N.; Sakthivel, P. Silver Decorated CeO2 Nanoparticles for Rapid Photocatalytic Degradation of Textile Rose Bengal Dye. Sci. Rep. 2021, 11, 1080. [Google Scholar] [CrossRef]
  95. Zhao, P.; Zhang, J.; Jiang, J.; Wang, H.; Xie, T.; Lin, Y. Synthesis and Study on Photogenerated Charge Behavior of Novel Pt/CeO2/ZnO Ternary Composites with Enhanced Photocatalytic Degradation Activity. J. Inorg. Organomet. Polym. Mater. 2020, 30, 1589–1599. [Google Scholar] [CrossRef]
  96. Killivalavan, G.; Sathyaseelan, B.; Kavitha, G.; Baskarann, I.; Senthilnathan, K.; Sivakumar, D.; Karthikeyan, N.; Manikandan, E.; Maaza, M. Cobalt Metal Ion Doped Cerium Oxide (Co-CeO2) Nanoparticles Effect Enhanced Photocatalytic Activity. MRS Adv. 2020, 5, 1–13. [Google Scholar] [CrossRef]
  97. Habib, I.Y.; Burhan, J.; Jaladi, F.; Lim, C.M.; Usman, A.; Kumara, N.T.R.N.; Tsang, S.C.E.; Mahadi, A.H. Effect of Cr Doping in CeO2 Nanostructures on Photocatalysis and H2O2 Assisted Methylene Blue Dye Degradation. Catal. Today 2021, 375, 506–513. [Google Scholar] [CrossRef]
  98. Majeed Khan, M.A.; Siwach, R.; Kumar, S.; Ahamed, M.; Ahmed, J. Frequency and Temperature Dependence of Dielectric Permittivity/Electric Modulus, and Efficient Photocatalytic Action of Fe-Doped CeO2 NPs. J. Alloys Compd. 2021, 856, 158127. [Google Scholar] [CrossRef]
  99. Majeed Khan, M.A.; Khan, W.; Naziruddin Khan, M.; Alhazaa, A.N. Enhanced Visible Light-Driven Photocatalytic Performance of Zr Doped CeO2 Nanoparticles. J. Mater. Sci. Mater. Electron. 2019, 30, 8291–8300. [Google Scholar] [CrossRef]
  100. Palanisamy, G.; Bhuvaneswari, K.; Bharathi, G.; Pazhanivel, T.; Grace, A.N.; Pasha, S.K.K. Construction of Magnetically Recoverable ZnS–WO3–CoFe2O4 Nanohybrid Enriched Photocatalyst for the Degradation of MB Dye under Visible Light Irradiation. Chemosphere 2021, 273, 129687. [Google Scholar] [CrossRef]
  101. Puga, F.; Navío, J.A.; Hidalgo, M.C. Features of Coupled AgBr/WO3 Materials as Potential Photocatalysts. J. Alloys Compd. 2021, 867, 159191. [Google Scholar] [CrossRef]
  102. Chen, X.; Li, Y.; Li, L. Facet-Engineered Surface and Interface Design of WO3/Bi2WO6 Photocatalyst with Direct Z-Scheme Heterojunction for Efficient Salicylic Acid Removal. Appl. Surf. Sci. 2020, 508, 144796. [Google Scholar] [CrossRef]
  103. Baral, B.; Mansingh, S.; Reddy, K.H.; Bariki, R.; Parida, K. Architecting a Double Charge-Transfer Dynamics In2S3/BiVO4 n-n Isotype Heterojunction for Superior Photocatalytic Oxytetracycline Hydrochloride Degradation and Water Oxidation Reaction: Unveiling the Association of Physicochemical, Electrochemical, and Ph. ACS Omega 2020, 5, 5270–5284. [Google Scholar] [CrossRef] [Green Version]
  104. Wang, W.; Han, Q.; Zhu, Z.; Zhang, L.; Zhong, S.; Liu, B. Enhanced Photocatalytic Degradation Performance of Organic Contaminants by Heterojunction Photocatalyst BiVO4/TiO2/RGO and Its Compatibility on Four Different Tetracycline Antibiotics. Adv. Powder Technol. 2019, 30, 1882–1896. [Google Scholar] [CrossRef]
  105. Guan, D.L.; Niu, C.G.; Wen, X.J.; Guo, H.; Deng, C.H.; Zeng, G.M. Enhanced Escherichia Coli Inactivation and Oxytetracycline Hydrochloride Degradation by a Z-Scheme Silver Iodide Decorated Bismuth Vanadate Nanocomposite under Visible Light Irradiation. J. Colloid Interface Sci. 2018, 512, 272–281. [Google Scholar] [CrossRef]
  106. Zhang, K.; Liu, Y.; Deng, J.; Xie, S.; Zhao, X.; Yang, J.; Han, Z.; Dai, H. Co–Pd/BiVO4: High-Performance Photocatalysts for the Degradation of Phenol under Visible Light Irradiation. Appl. Catal. B Environ. 2018, 224, 350–359. [Google Scholar] [CrossRef]
  107. Deng, Y.; Tang, L.; Zeng, G.; Wang, J.; Zhou, Y.; Wang, J.; Tang, J.; Wang, L.; Feng, C. Facile Fabrication of Mediator-Free Z-Scheme Photocatalyst of Phosphorous-Doped Ultrathin Graphitic Carbon Nitride Nanosheets and Bismuth Vanadate Composites with Enhanced Tetracycline Degradation under Visible Light. J. Colloid Interface Sci. 2018, 509, 219–234. [Google Scholar] [CrossRef] [PubMed]
  108. Samsudin, M.F.R.; Mahmood, A.; Sufian, S. Enhanced Photocatalytic Degradation of Wastewater over RGO-TiO2/BiVO4 Photocatalyst under Solar Light Irradiation. J. Mol. Liq. 2018, 268, 26–36. [Google Scholar] [CrossRef]
  109. Wang, J.; Jin, J.; Wang, X.; Yang, S.; Zhao, Y.; Wu, Y.; Dong, S.; Sun, J.; Sun, J. Facile Fabrication of Novel BiVO4/Bi2S3/MoS2 n-p Heterojunction with Enhanced Photocatalytic Activities towards Pollutant Degradation under Natural Sunlight. J. Colloid Interface Sci. 2017, 505, 805–815. [Google Scholar] [CrossRef] [PubMed]
  110. Ma, D.K.; Guan, M.L.; Liu, S.S.; Zhang, Y.Q.; Zhang, C.W.; He, Y.X.; Huang, S.M. Controlled Synthesis of Olive-Shaped Bi2S3/BiVO4 Microspheres through a Limited Chemical Conversion Route and Enhanced Visible-Light-Responding Photocatalytic Activity. Dalton Trans. 2012, 41, 5581–5586. [Google Scholar] [CrossRef]
  111. Wei, Z.; Benlin, D.; Fengxia, Z.; Xinyue, T.; Jiming, X.; Lili, Z.; Shiyin, L.; Leung, D.Y.C.; Sun, C. A Novel 3D Plasmonic P-n Heterojunction Photocatalyst: Ag Nanoparticles on Flower-like p-Ag2S/n-BiVO4 and Its Excellent Photocatalytic Reduction and Oxidation Activities. Appl. Catal. B Environ. 2018, 229, 171–180. [Google Scholar] [CrossRef]
  112. Sreethawong, T.; Laehsalee, S.; Chauadej, S. Comparative Investigation of Mesoporous- and Non-Mesoporous-Assembled TiO2 Nanocrystals for Photocatalytic H2 Production over N-Doped TiO2 under Visible Light Irradiation. Int. J. Hydrogen Energy 2008, 33, 5947–5957. [Google Scholar] [CrossRef]
  113. Lin, T.; Yang, C.; Wang, Z.; Yin, H.; Lü, X.; Huang, F.; Lin, J.; Xie, X.; Jiang, M. Effective Nonmetal Incorporation in Black Titania with Enhanced Solar Energy Utilization. Energy Environ. Sci. 2014, 7, 967–972. [Google Scholar] [CrossRef]
  114. Villa, K.; Black, A.; Domènech, X.; Peral, J. Nitrogen Doped TiO2 for Hydrogen Production under Visible Light Irradiation. Sol. Energy 2012, 86, 558–566. [Google Scholar] [CrossRef]
  115. Wu, F.; Hu, X.; Fan, J.; Liu, E.; Sun, T.; Kang, L.; Hou, W.; Zhu, C.; Liu, H. Photocatalytic Activity of Ag/TiO2 Nanotube Arrays Enhanced by Surface Plasmon Resonance and Application in Hydrogen Evolution by Water Splitting. Plasmonics 2013, 8, 501–508. [Google Scholar] [CrossRef]
  116. Fan, X.; Wan, J.; Liu, E.; Sun, L.; Hu, Y.; Li, H.; Hu, X.; Fan, J. High-Efficiency Photoelectrocatalytic Hydrogen Generation Enabled by Ag Deposited and Ce Doped TiO2 Nanotube Arrays. Ceram. Int. 2015, 41, 5107–5116. [Google Scholar] [CrossRef]
  117. Chen, W.T.; Chan, A.; Sun-Waterhouse, D.; Moriga, T.; Idriss, H.; Waterhouse, G.I.N. Ni/TiO2: A Promising Low-Cost Photocatalytic System for Solar H2 Production from Ethanol-Water Mixtures. J. Catal. 2015, 326, 43–53. [Google Scholar] [CrossRef]
  118. Dholam, R.; Patel, N.; Adami, M.; Miotello, A. Hydrogen Production by Photocatalytic Water-Splitting Using Cr- or Fe-Doped TiO2 Composite Thin Films Photocatalyst. Int. J. Hydrogen Energy 2009, 34, 5337–5346. [Google Scholar] [CrossRef]
  119. Sadanandam, G.; Lalitha, K.; Kumari, V.D.; Shankar, M.V.; Subrahmanyam, M. Cobalt Doped TiO2: A Stable and Efficient Photocatalyst for Continuous Hydrogen Production from Glycerol: Water Mixtures under Solar Light Irradiation. Int. J. Hydrogen Energy 2013, 38, 9655–9664. [Google Scholar] [CrossRef]
  120. Tentu, R.D.; Basu, S. Photocatalytic Water Splitting for Hydrogen Production. Curr. Opin. Electrochem. 2017, 5, 56–62. [Google Scholar] [CrossRef]
  121. De Lasa, H.; Rosales, B.S.; Moreira, J.; Valades-Pelayo, P. Efficiency Factors in Photocatalytic Reactors: Quantum Yield and Photochemical Thermodynamic Efficiency Factor. Chem. Eng. Technol. 2016, 39, 51–65. [Google Scholar] [CrossRef]
  122. Corcoran, E.B.; McMullen, J.P.; Lévesque, F.; Wismer, M.K.; Naber, J.R. Photon Equivalents as a Parameter for Scaling Photoredox Reactions in Flow: Translation of Photocatalytic C−N Cross-Coupling from Lab Scale to Multikilogram Scale. Angew. Chem. Int. Ed. 2020, 59, 11964–11968. [Google Scholar] [CrossRef]
  123. Azzaz, A.A.; Assadi, A.A.; Jellali, S.; Bouzaza, A.; Wolbert, D.; Rtimi, S.; Bousselmi, L. Discoloration of Simulated Textile Effluent in Continuous Photoreactor Using Immobilized Titanium Dioxide: Effect of Zinc and Sodium Chloride. J. Photochem. Photobiol. A Chem. 2018, 358, 111–120. [Google Scholar] [CrossRef]
  124. Pedanekar, R.S.; Shaikh, S.K.; Rajpure, K.Y. Thin Film Photocatalysis for Environmental Remediation: A Status Review. Curr. Appl. Phys. 2020, 20, 931–952. [Google Scholar] [CrossRef]
  125. Montecchio, F.; Chinungi, D.; Lanza, R.; Engvall, K. Surface Treatments of Metal Supports for Photocatalysis Applications. Appl. Surf. Sci. 2017, 401, 283–296. [Google Scholar] [CrossRef]
  126. Mohd Adnan, M.A.; Muhd Julkapli, N.; Amir, M.N.I.; Maamor, A. Effect on Different TiO2 Photocatalyst Supports on Photodecolorization of Synthetic Dyes: A Review. Int. J. Environ. Sci. Technol. 2019, 16, 547–566. [Google Scholar] [CrossRef]
  127. Shi, E.; Xu, Z.; Wang, W.; Xu, Y.; Zhang, Y.; Yang, X.; Liu, Q.; Zeng, T.; Song, S.; Jiang, Y.; et al. Ag2S-Doped Core-Shell Nanostructures of Fe3O4@Ag3PO4 Ultrathin Film: Major Role of Hole in Rapid Degradation of Pollutants under Visible Light Irradiation. Chem. Eng. J. 2019, 366, 123–132. [Google Scholar] [CrossRef]
  128. Abdel-Maksoud, Y.; Imam, E.; Ramadan, A. TiO2 Solar Photocatalytic Reactor Systems: Selection of Reactor Design for Scale-up and Commercialization—Analytical Review. Catalysts 2016, 6, 138. [Google Scholar] [CrossRef] [Green Version]
  129. Toe, C.Y.; Pan, J.; Scott, J.; Amal, R. Identifying Key Design Criteria for Large-Scale Photocatalytic Hydrogen Generation from Engineering and Economic Perspectives. ACS ES&T Eng. 2022, 2, 1130–1143. [Google Scholar] [CrossRef]
  130. Visan, A.; Van Ommen, J.R.; Kreutzer, M.T.; Lammertink, R.G.H. Photocatalytic Reactor Design: Guidelines for Kinetic Investigation. Ind. Eng. Chem. Res. 2019, 58, 5349–5357. [Google Scholar] [CrossRef] [Green Version]
  131. Wetchakun, K.; Wetchakun, N.; Sakulsermsuk, S. An Overview of Solar/Visible Light-Driven Heterogeneous Photocatalysis for Water Purification: TiO2- and ZnO-Based Photocatalysts Used in Suspension Photoreactors. J. Ind. Eng. Chem. 2019, 71, 19–49. [Google Scholar] [CrossRef]
  132. Wood, D.; Shaw, S.; Cawte, T.; Shanen, E.; Van Heyst, B. An Overview of Photocatalyst Immobilization Methods for Air Pollution Remediation. Chem. Eng. J. 2020, 391, 123490. [Google Scholar] [CrossRef]
  133. Uekert, T.; Pichler, C.M.; Schubert, T.; Reisner, E. Solar-Driven Reforming of Solid Waste for a Sustainable Future. Nat. Sustain. 2021, 4, 389–391. [Google Scholar] [CrossRef]
  134. Sundar, K.P.; Kanmani, S. Progression of Photocatalytic Reactors and It’s Comparison: A Review. Chem. Eng. Res. Des. 2020, 154, 135–150. [Google Scholar] [CrossRef]
  135. Azad Gilani, H.; Hoseinzadeh, S. Techno-Economic Study of Compound Parabolic Collector in Solar Water Heating System in the Northern Hemisphere. Appl. Therm. Eng. 2021, 190, 116756. [Google Scholar] [CrossRef]
  136. Chong, M.N.; Jin, B.; Chow, C.W.K.; Saint, C. Recent Developments in Photocatalytic Water Treatment Technology: A Review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef]
  137. Cabrera Reina, A.; Miralles-Cuevas, S.; Cornejo, L.; Pomares, L.; Polo, J.; Oller, I.; Malato, S. The Influence of Location on Solar Photo-Fenton: Process Performance, Photoreactor Scaling-up and Treatment Cost. Renew. Energy 2020, 145, 1890–1900. [Google Scholar] [CrossRef]
  138. Rodríguez, S.M.; Gálvez, J.B.; Rubio, M.I.M.; Ibáñez, P.F.; Padilla, D.A.; Pereira, M.C.; Mendes, J.F.; De Oliveira, J.C. Engineering of Solar Photocatalytic Collectors. Sol. Energy 2004, 77, 513–524. [Google Scholar] [CrossRef]
  139. Spasiano, D.; Marotta, R.; Malato, S.; Fernandez-Ibañez, P.; Di Somma, I. Solar Photocatalysis: Materials, Reactors, Some Commercial, and Pre-Industrialized Applications. A Comprehensive Approach. Appl. Catal. B Environ. 2015, 170–171, 90–123. [Google Scholar] [CrossRef]
  140. Bahnemann, D. Photocatalytic Water Treatment: Solar Energy Applications. Sol. Energy 2004, 77, 445–459. [Google Scholar] [CrossRef]
  141. Malato, S.; Blanco, J.; Alarcón, D.C.; Maldonado, M.I.; Fernández-Ibáñez, P.; Gernjak, W. Photocatalytic Decontamination and Disinfection of Water with Solar Collectors. Catal. Today 2007, 122, 137–149. [Google Scholar] [CrossRef]
  142. Fernández-Pérez, A.; Rodríguez-Casado, V.; Valdés-Solís, T.; Marbán, G. A New Continuous Flow-through Structured Reactor for the Photodegradation of Aqueous Contaminants. J. Environ. Chem. Eng. 2018, 6, 4070–4077. [Google Scholar] [CrossRef]
  143. Nawaz, A.; Kuila, A.; Rani, A.; Mishra, N.S.; Sim, L.C.; Leong, K.H.; Saravanan, P. Industrial Application of Light-Driven Nanomaterial. In Industrial Applications of Nanomaterials; Elsevier Inc.: Amsterdam, The Netherlands, 2019; pp. 151–179. ISBN 9780128157497. [Google Scholar]
  144. SOWARLA. Demonstration Plant. Available online: http://www.sowarla.de/demonstration-plant.html (accessed on 28 September 2022).
  145. Sattler, C.; Jung, C.; Senholdt, M.; Nietsch, A.; Olwig, R.; Hulth, R. Solar Applications in Water Treatment; SOWARLA GmbH: Eberstadt, Germany, 2012; pp. 1–25. [Google Scholar]
  146. Sacco, O.; Vaiano, V.; Sannino, D. Main Parameters Influencing the Design of Photocatalytic Reactors for Wastewater Treatment: A Mini Review. J. Chem. Technol. Biotechnol. 2020, 95, 2608–2618. [Google Scholar] [CrossRef]
  147. Benotti, M.J.; Stanford, B.D.; Wert, E.C.; Snyder, S.A. Evaluation of a Photocatalytic Reactor Membrane Pilot System for the Removal of Pharmaceuticals and Endocrine Disrupting Compounds from Water. Water Res. 2009, 43, 1513–1522. [Google Scholar] [CrossRef]
  148. Mozia, S. Photocatalytic Membrane Reactors (PMRs) in Water and Wastewater Treatment. A Review. Sep. Purif. Technol. 2010, 73, 71–91. [Google Scholar] [CrossRef]
  149. Sayama, K.; Nomura, A.; Arai, T.; Sugita, T.; Abe, R.; Yanagida, M.; Oi, T.; Iwasaki, Y.; Abe, Y.; Sugihara, H. Photoelectrochemical Decomposition of Water into H2 and O2 on Porous BiVO4 Thin-Film Electrodes under Visible Light and Significant Effect of Ag Ion Treatment. J. Phys. Chem. B 2006, 110, 11352–11360. [Google Scholar] [CrossRef] [PubMed]
  150. Nomura, Y.; Fukahori, S.; Fujiwara, T. Removal of Sulfamonomethoxine and Its Transformation Byproducts from Fresh Aquaculture Wastewater by a Rotating Advanced Oxidation Contactor Equipped with Zeolite/TiO2 Composite Sheets. Process Saf. Environ. Prot. 2020, 134, 161–168. [Google Scholar] [CrossRef]
  151. Bian, Z.; Cao, F.; Zhu, J.; Li, H. Plant Uptake-Assisted Round-the-Clock Photocatalysis for Complete Purification of Aquaculture Wastewater Using Sunlight. Environ. Sci. Technol. 2015, 49, 2418–2424. [Google Scholar] [CrossRef] [PubMed]
  152. Khan, S.J.; Reed, R.H.; Rasul, M.G. Thin-Film Fixed-Bed Reactor for Solar Photocatalytic Inactivation of Aeromonas Hydrophila: Influence of Water Quality. BMC Microbiol. 2012, 12, 285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Bousselmi, L.; Geissen, S.U.; Schroeder, H. Textile Wastewater Treatment and Reuse by Solar Catalysis: Results from a Pilot Plant in Tunisia. Water Sci. Technol. 2004, 49, 331–337. [Google Scholar] [CrossRef]
  154. Isawi, H.; Abdelaziz, M.O.; Abo Zeed, D.; El-Kholy, R.A.; El-Noss, M.; Said, M.M.; El-Aassar, A.H.M.; Shawky, H.A. Semi Industrial Continuous Flow Photoreactor for Wastewater Purification in Some Polluted Areas: Design, Manufacture, and Socio-Economic Impacts. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100544. [Google Scholar] [CrossRef]
  155. EL-Mekkawi, D.M.; Abdelwahab, N.A.; Mohamed, W.A.A.; Taha, N.A.; Abdel-Mottaleb, M.S.A. Solar Photocatalytic Treatment of Industrial Wastewater Utilizing Recycled Polymeric Disposals as TiO2 Supports. J. Clean. Prod. 2020, 249, 119430. [Google Scholar] [CrossRef]
  156. Schröder, M.; Kailasam, K.; Borgmeyer, J.; Neumann, M.; Thomas, A.; Schomäcker, R.; Schwarze, M. Hydrogen Evolution Reaction in a Large-Scale Reactor Using a Carbon Nitride Photocatalyst under Natural Sunlight Irradiation. Energy Technol. 2015, 3, 1014–1017. [Google Scholar] [CrossRef]
  157. Maldonado, M.I.; López-Martín, A.; Colón, G.; Peral, J.; Martínez-Costa, J.I.; Malato, S. Solar Pilot Plant Scale Hydrogen Generation by Irradiation of Cu/TiO2 Composites in Presence of Sacrificial Electron Donors. Appl. Catal. B Environ. 2018, 229, 15–23. [Google Scholar] [CrossRef]
  158. Villa, K.; Domènech, X.; Malato, S.; Maldonado, M.I.; Peral, J. Heterogeneous Photocatalytic Hydrogen Generation in a Solar Pilot Plant. Int. J. Hydrogen Energy 2013, 38, 12718–12724. [Google Scholar] [CrossRef]
  159. Almomani, F.; Shawaqfah, M.; Alkasrawi, M. Solar-Driven Hydrogen Production from a Water-Splitting Cycle Based on Carbon-TiO2 Nano-Tubes. Int. J. Hydrogen Energy 2022, 47, 3294–3305. [Google Scholar] [CrossRef]
  160. Ren, Y.; Zhao, L.; Jing, D.; Guo, L. Investigation and Modeling of CPC Based Tubular Photocatalytic Reactor for Scaled-up Hydrogen Production. Int. J. Hydrogen Energy 2016, 41, 16019–16031. [Google Scholar] [CrossRef]
  161. Arzate Salgado, S.Y.; Ramírez Zamora, R.M.; Zanella, R.; Peral, J.; Malato, S.; Maldonado, M.I. Photocatalytic Hydrogen Production in a Solar Pilot Plant Using a Au/TiO2 Photo Catalyst. Int. J. Hydrogen Energy 2016, 41, 11933–11940. [Google Scholar] [CrossRef]
  162. Liu, X.; Cao, Z.; Yuan, Z.; Zhang, J.; Guo, X.; Yang, Y.; He, F.; Zhao, Y.; Xu, J. Insight into the Kinetics and Mechanism of Removal of Aqueous Chlorinated Nitroaromatic Antibiotic Chloramphenicol by Nanoscale Zero-Valent Iron. Chem. Eng. J. 2018, 334, 508–518. [Google Scholar] [CrossRef]
  163. Nishiyama, H.; Yamada, T.; Nakabayashi, M.; Maehara, Y.; Yamaguchi, M.; Kuromiya, Y.; Nagatsuma, Y.; Tokudome, H.; Akiyama, S.; Watanabe, T.; et al. Photocatalytic Solar Hydrogen Production from Water on a 100-M2 Scale. Nature 2021, 598, 304–307. [Google Scholar] [CrossRef]
  164. Oros-Ruiz, S.; Zanella, R.; Collins, S.E.; Hernández-Gordillo, A.; Gómez, R. Photocatalytic Hydrogen Production by Au-MxOy (MAg, Cu, Ni) Catalysts Supported on TiO2. Catal. Commun. 2014, 47, 1–6. [Google Scholar] [CrossRef]
  165. Fang, W.; Shangguan, W. A Review on Bismuth-Based Composite Oxides for Photocatalytic Hydrogen Generation. Int. J. Hydrogen Energy 2019, 44, 895–912. [Google Scholar] [CrossRef]
  166. Zhu, Z.; Wan, S.; Zhao, Y.; Gu, Y.; Wang, Y.; Qin, Y.; Zhang, Z.; Ge, X.; Zhong, Q.; Bu, Y. Recent Advances in Bismuth-Based Multimetal Oxide Photocatalysts for Hydrogen Production from Water Splitting: Competitiveness, Challenges, and Future Perspectives. Mater. Rep. Energy 2021, 1, 100019. [Google Scholar] [CrossRef]
  167. Asiri, A.M.; Nawaz, T.; Tahir, M.B.; Fatima, N.; Khan, S.B.; Alamry, K.A.; Alfifi, S.Y.; Marwani, H.M.; Al-Otaibi, M.M.; Chakraborty, S. Fabrication of WO3 Based Nanocomposites for the Excellent Photocatalytic Energy Production under Visible Light Irradiation. Int. J. Hydrogen Energy 2021, 46, 39058–39066. [Google Scholar] [CrossRef]
  168. Ning, X.; Lu, G. Photocorrosion Inhibition of CdS-Based Catalysts for Photocatalytic Overall Water Splitting. Nanoscale 2020, 12, 1213–1223. [Google Scholar] [CrossRef]
  169. Younis, S.A.; Kim, K.H. Heterogeneous Photocatalysis Scalability for Environmental Remediation: Opportunities and Challenges. Catalysts 2020, 10, 1109. [Google Scholar] [CrossRef]
  170. Dwivedi, S.K.; Vishwakarma, M. Hydrogen Embrittlement in Different Materials: A Review. Int. J. Hydrogen Energy 2018, 43, 21603–21616. [Google Scholar] [CrossRef]
  171. Kim, E.; Park, J.; Cho, J.H.; Moon, I. Simulation of Hydrogen Leak and Explosion for the Safety Design of Hydrogen Fueling Station in Korea. Int. J. Hydrogen Energy 2013, 38, 1737–1743. [Google Scholar] [CrossRef]
  172. Abohamzeh, E.; Salehi, F.; Sheikholeslami, M.; Abbassi, R.; Khan, F. Review of Hydrogen Safety during Storage, Transmission, and Applications Processes. J. Loss Prev. Process Ind. 2021, 72, 104569. [Google Scholar] [CrossRef]
  173. Durbin, D.J.; Malardier-Jugroot, C. Review of Hydrogen Storage Techniques for on Board Vehicle Applications. Int. J. Hydrogen Energy 2013, 38, 14595–14617. [Google Scholar] [CrossRef]
  174. Moradi, R.; Groth, K.M. Hydrogen Storage and Delivery: Review of the State of the Art Technologies and Risk and Reliability Analysis. Int. J. Hydrogen Energy 2019, 44, 12254–12269. [Google Scholar] [CrossRef]
  175. Nof, S.Y. Automation: What It Means to Us around the World. In Springer Handbook of Automation; Springer: Berlin/Heidelberg, Germany, 2009; pp. 13–52. [Google Scholar]
  176. Bai, Y.; Wilbraham, L.; Slater, B.J.; Zwijnenburg, M.A.; Sprick, R.S.; Cooper, A.I. Accelerated Discovery of Organic Polymer Photocatalysts for Hydrogen Evolution from Water through the Integration of Experiment and Theory. J. Am. Chem. Soc. 2019, 141, 9063–9071. [Google Scholar] [CrossRef] [Green Version]
  177. Lopato, E.M.; Bernhard, S. Exploring Multidimensional Chemical Spaces: Instrumentation and Chemical Systems for the Parallelization of Hydrogen Evolving Photocatalytic Reactions. Energy Fuels 2021, 35, 18957–18981. [Google Scholar] [CrossRef]
  178. Zhao, C.; Krall, A.; Zhao, H.; Zhang, Q.; Li, Y. Ultrasonic Spray Pyrolysis Synthesis of Ag/TiO2 Nanocomposite Photocatalysts for Simultaneous H2 Production and CO2 Reduction. Int. J. Hydrogen Energy 2012, 37, 9967–9976. [Google Scholar] [CrossRef]
  179. Anjum Moinuddin, A.; Vijay Kotkondawar, A.; Hippargi, G.; Anshul, A.; Rayalu, S. Morphologically and Hierarchically Controlled Ag/Ag2MoO4 Microspheres for Photocatalytic Hydrogen Generation. Appl. Surf. Sci. 2022, 597, 153554. [Google Scholar] [CrossRef]
  180. Li, J.; Li, J.; Liu, R.; Tu, Y.; Li, Y.; Cheng, J.; He, T.; Zhu, X. Autonomous Discovery of Optically Active Chiral Inorganic Perovskite Nanocrystals through an Intelligent Cloud Lab. Nat. Commun. 2020, 11, 2046. [Google Scholar] [CrossRef]
  181. Wang, J.; Evans, J.R.G. London University Search Instrument: A Combinatorial Robot for High-Throughput Methods in Ceramic Science. J. Comb. Chem. 2005, 7, 665–672. [Google Scholar] [CrossRef]
  182. Burger, B.; Maffettone, P.M.; Gusev, V.V.; Aitchison, C.M.; Bai, Y.; Wang, X.; Li, X.; Alston, B.M.; Li, B.; Clowes, R.; et al. A Mobile Robotic Chemist. Nature 2020, 583, 237–241. [Google Scholar] [CrossRef]
  183. MacLeod, B.P.; Parlane, F.G.L.; Brown, A.K.; Hein, J.E.; Berlinguette, C.P. Flexible Automation Accelerates Materials Discovery. Nat. Mater. 2022, 21, 722–726. [Google Scholar] [CrossRef]
  184. Sa, B.; Hu, R.; Zheng, Z.; Xiong, R.; Zhang, Y.; Wen, C.; Zhou, J.; Sun, Z. High-Throughput Computational Screening and Machine Learning Modeling of Janus 2D III-VI van Der Waals Heterostructures for Solar Energy Applications. Chem. Mater. 2022, 34, 6687–6701. [Google Scholar] [CrossRef]
  185. Wildey, M.J.; Haunso, A.; Tudor, M.; Webb, M.; Connick, J.H. Chapter Five—High-Throughput Screening. Annu. Rep. Med. Chem. 2017, 50, 149–195. [Google Scholar] [CrossRef]
  186. Moradi, S.; Kundu, S.; Saidaminov, M.I. High-Throughput Synthesis of Thin Films for the Discovery of Energy Materials: A Perspective. ACS Mater. Au 2022, 2, 516–524. [Google Scholar] [CrossRef]
  187. Nursam, N.M.; Wang, X.; Caruso, R.A. High-Throughput Synthesis and Screening of Titania-Based Photocatalysts. ACS Comb. Sci. 2015, 17, 548–569. [Google Scholar] [CrossRef]
  188. Sun, S.; Ding, J.; Bao, J.; Luo, Z.; Gao, C. Application of Parallel Synthesis and High Throughput Characterization in Photocatalyst Discovery. Comb. Chem. High Throughput Screen. 2011, 14, 160–172. [Google Scholar] [CrossRef]
  189. Lewis, N.S.; Katz, J.; Gingrish, T. High-Throughput Screening and Device for Photocatalysts; California Institute of Technology (CalTech): Pasadena, CA, USA, 2010; pp. 1–11. [Google Scholar]
  190. Kumari, S.; Gutkowski, R.; Junqueira, J.R.C.; Kostka, A.; Hengge, K.; Scheu, C.; Schuhmann, W.; Ludwig, A. Combinatorial Synthesis and High-Throughput Characterization of Fe-V-O Thin-Film Materials Libraries for Solar Water Splitting. ACS Comb. Sci. 2018, 20, 544–553. [Google Scholar] [CrossRef]
  191. Kumari, S.; Junqueira, J.R.C.; Schuhmann, W.; Ludwig, A. High-Throughput Exploration of Metal Vanadate Thin-Film Systems (M-V-O, M = Cu, Ag, W, Cr, Co, Fe) for Solar Water Splitting: Composition, Structure, Stability, and Photoelectrochemical Properties. ACS Comb. Sci. 2020, 22, 844–857. [Google Scholar] [CrossRef] [PubMed]
  192. Stroyuk, O.; Raievska, O.; Langner, S.; Kupfer, C.; Barabash, A.; Solonenko, D.; Azhniuk, Y.; Hauch, J.; Osvet, A.; Batentschuk, M.; et al. High-Throughput Robotic Synthesis and Photoluminescence Characterization of Aqueous Multinary Copper–Silver Indium Chalcogenide Quantum Dots. Part. Part. Syst. Charact. 2021, 38, 2100169. [Google Scholar] [CrossRef]
  193. Hattrick-Simpers, J.R.; Zakutayev, A.; Barron, S.C.; Trautt, Z.T.; Nguyen, N.; Choudhary, K.; Decost, B.; Phillips, C.; Kusne, A.G.; Yi, F.; et al. An Inter-Laboratory Study of Zn-Sn-Ti-O Thin Films Using High-Throughput Experimental Methods. ACS Comb. Sci. 2019, 21, 350–361. [Google Scholar] [CrossRef] [PubMed]
  194. Visan, A.; Rafieian, D.; Ogieglo, W.; Lammertink, R.G.H. Modeling Intrinsic Kinetics in Immobilized Photocatalytic Microreactors. Appl. Catal. B Environ. 2014, 150–151, 93–100. [Google Scholar] [CrossRef]
  195. Claes, T.; Dilissen, A.; Leblebici, M.E.; Van Gerven, T. Translucent Packed Bed Structures for High Throughput Photocatalytic Reactors. Chem. Eng. J. 2019, 361, 725–735. [Google Scholar] [CrossRef]
  196. Manassero, A.; Satuf, M.L.; Alfano, O.M. Photocatalytic Reactors with Suspended and Immobilized TiO2: Comparative Efficiency Evaluation. Chem. Eng. J. 2017, 326, 29–36. [Google Scholar] [CrossRef]
  197. Gao, Z.; Pan, C.; Choi, C.; Chang, C. Continuous-Flow Photocatalytic Microfluidic-Reactor for the Treatment of Aqueous Contaminants, Simplicity, and Complexity: A Mini-Review. Symmetry 2021, 13, 1325. [Google Scholar] [CrossRef]
  198. De Sá, D.S.; Vasconcelos, L.E.; de Souza, J.R.; Marinkovic, B.A.; Del Rosso, T.; Fulvio, D.; Maza, D.; Massi, A.; Pandoli, O. Intensification of Photocatalytic Degradation of Organic Dyes and Phenol by Scale-up and Numbering-up of Meso- and Microfluidic TiO2 Reactors for Wastewater Treatment. J. Photochem. Photobiol. A Chem. 2018, 364, 59–75. [Google Scholar] [CrossRef]
  199. Hamaloğlu, K.Ö.; Sağ, E.; Tuncel, A. Bare, Gold and Silver Nanoparticle Decorated, Monodisperse-Porous Titania Microbeads for Photocatalytic Dye Degradation in a Newly Constructed Microfluidic, Photocatalytic Packed-Bed Reactor. J. Photochem. Photobiol. A Chem. 2017, 332, 60–65. [Google Scholar] [CrossRef]
  200. Zhao, P.; Qin, N.; Wen, J.Z.; Ren, C.L. Photocatalytic Performances of ZnO Nanoparticle Film and Vertically Aligned Nanorods in Chamber-Based Microfluidic Reactors: Reaction Kinetics and Flow Effects. Appl. Catal. B Environ. 2017, 209, 468–475. [Google Scholar] [CrossRef]
  201. Gao, W.; Jing, W.; Li, Z.; Wu, Q.; Han, F.; Zhao, L.; Yang, Z.; Jiang, Z. Facile Fabrication of Crescentic ZnO Nanorod-Based Photo-Catalytic Micro-Fluidic Reactors. Microelectron. Eng. 2022, 263, 111843. [Google Scholar] [CrossRef]
  202. Qin, S.; Lei, Y.; Huang, J.F.; Xiao, L.M.; Liu, J.M. A Robust Photocatalytic Hybrid Material Composed of Metal-Organic Cages and TiO2 for Efficient Visible-Light-Driven Hydrogen Evolution. Chem. Asian J. 2021, 16, 2055–2062. [Google Scholar] [CrossRef]
  203. Liang, L.; Shi, L.; Wang, F.; Wang, H.; Yan, P.; Cong, Y.; Yao, L.; Yang, Z.; Qi, W. G-C3N4 Nano-Fragments as Highly Efficient Hydrogen Evolution Photocatalysts: Boosting Effect of Nitrogen Vacancy. Appl. Catal. A Gen. 2020, 599, 117618. [Google Scholar] [CrossRef]
  204. Guo, F.; Wang, L.; Sun, H.; Li, M.; Shi, W.; Lin, X. A One-Pot Sealed Ammonia Self-Etching Strategy to Synthesis of N-Defective g-C3N4 for Enhanced Visible-Light Photocatalytic Hydrogen. Int. J. Hydrogen Energy 2020, 45, 30521–30532. [Google Scholar] [CrossRef]
  205. Ludwig, A. Discovery of New Materials Using Combinatorial Synthesis and High-Throughput Characterization of Thin-Film Materials Libraries Combined with Computational Methods. npj Comput. Mater. 2019, 70, 1038. [Google Scholar] [CrossRef] [Green Version]
  206. Luo, S.; Li, T.; Wang, X.; Faizan, M.; Zhang, L. High-Throughput Computational Materials Screening and Discovery of Optoelectronic Semiconductors. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2021, 11, e1489. [Google Scholar] [CrossRef]
  207. Ren, E.; Guilbaud, P.; Coudert, F.-X. High-Throughput Computational Screening of Nanoporous Materials in Targeted Applications. Digit. Discov. 2022, 1, 355–374. [Google Scholar] [CrossRef]
  208. Zhang, X.; Zhang, Z.; Wu, D.; Zhang, X.; Zhao, X.; Zhou, Z. Computational Screening of 2D Materials and Rational Design of Heterojunctions for Water Splitting Photocatalysts. Small Methods 2018, 2, 1700359. [Google Scholar] [CrossRef]
  209. Wu, Y.; Lazic, P.; Hautier, G.; Persson, K.; Ceder, G. First Principles High Throughput Screening of Oxynitrides for Water-Splitting Photocatalysts. Energy Environ. Sci. 2013, 6, 157–168. [Google Scholar] [CrossRef] [Green Version]
  210. Lin, C.; Feng, X.; Legut, D.; Liu, X.; Seh, Z.W.; Zhang, R.; Zhang, Q. Discovery of Efficient Visible-Light Driven Oxygen Evolution Photocatalysts: Automated High-Throughput Computational Screening of MA2Z4. Adv. Funct. Mater. 2022, 32, 2207415. [Google Scholar] [CrossRef]
  211. Singh, A.K.; Mathew, K.; Zhuang, H.L.; Hennig, R.G. Computational Screening of 2D Materials for Photocatalysis. J. Phys. Chem. Lett. 2015, 6, 1087–1098. [Google Scholar] [CrossRef] [PubMed]
  212. Liu, Z.; Wang, B.; Chu, D.; Cazorla, C. First-Principles High-Throughput Screening of Bulk Piezo-Photocatalytic Materials for Sunlight-Driven Hydrogen Production. J. Mater. Chem. A 2022, 10, 18132–18146. [Google Scholar] [CrossRef]
  213. Mai, H.; Le, T.C.; Chen, D.; Winkler, D.A.; Caruso, R.A. Machine Learning for Electrocatalyst and Photocatalyst Design and Discovery. Chem. Rev. 2022, 122, 13478–13515. [Google Scholar] [CrossRef] [PubMed]
  214. Masood, H.; Toe, C.Y.; Teoh, W.Y.; Sethu, V.; Amal, R. Machine Learning for Accelerated Discovery of Solar Photocatalysts. ACS Catal. 2019, 9, 11774–11787. [Google Scholar] [CrossRef]
  215. Tao, H.; Wu, T.; Kheiri, S.; Aldeghi, M.; Aspuru-Guzik, A.; Kumacheva, E. Self-Driving Platform for Metal Nanoparticle Synthesis: Combining Microfluidics and Machine Learning. Adv. Funct. Mater. 2021, 31, 2106725. [Google Scholar] [CrossRef]
Figure 1. Heterogeneous photocatalytic process to degrade organic pollutants in water and green H2 production through the water splitting reaction.
Figure 1. Heterogeneous photocatalytic process to degrade organic pollutants in water and green H2 production through the water splitting reaction.
Catalysts 13 00154 g001
Figure 2. Comparison of the number of published studies on the synthesis of new photocatalysts (green bars), the development of photoreactors (blue bars), and the upscaling of the heterogeneous photocatalysis process (yellow bars). The search of the publication was performed in Scopus using the keywords “photocatalysis” + “synthesis”, “photocatalysis” + “reactors”, and “photocatalysis” + “reactors” + “scaling”.
Figure 2. Comparison of the number of published studies on the synthesis of new photocatalysts (green bars), the development of photoreactors (blue bars), and the upscaling of the heterogeneous photocatalysis process (yellow bars). The search of the publication was performed in Scopus using the keywords “photocatalysis” + “synthesis”, “photocatalysis” + “reactors”, and “photocatalysis” + “reactors” + “scaling”.
Catalysts 13 00154 g002
Figure 3. The Green Chemistry Principles (GCP, with information from [30]).
Figure 3. The Green Chemistry Principles (GCP, with information from [30]).
Catalysts 13 00154 g003
Figure 4. Green Engineering Principles (GEP, with information of [26]).
Figure 4. Green Engineering Principles (GEP, with information of [26]).
Catalysts 13 00154 g004
Figure 5. Periodic Table’s endangered elements.
Figure 5. Periodic Table’s endangered elements.
Catalysts 13 00154 g005
Figure 6. Global horizontal irradiation (Global Solar Atlas).
Figure 6. Global horizontal irradiation (Global Solar Atlas).
Catalysts 13 00154 g006
Figure 7. Non-concentrating solar collector tested at the Plataforma Solar de Almería, Spain (reproduced with permission of [141]).
Figure 7. Non-concentrating solar collector tested at the Plataforma Solar de Almería, Spain (reproduced with permission of [141]).
Catalysts 13 00154 g007
Figure 8. The SOWARLA demonstration plant, (taken with permission from [143]).
Figure 8. The SOWARLA demonstration plant, (taken with permission from [143]).
Catalysts 13 00154 g008
Figure 9. Rotating advanced oxidation contactor (RAOC) equipped with zeolite/TiO2 composite sheets (reproduced with permission of [150]).
Figure 9. Rotating advanced oxidation contactor (RAOC) equipped with zeolite/TiO2 composite sheets (reproduced with permission of [150]).
Catalysts 13 00154 g009
Figure 10. Chemist robot at laboratory scale used for the synthesis of photocatalysts to produce green H2 [182].
Figure 10. Chemist robot at laboratory scale used for the synthesis of photocatalysts to produce green H2 [182].
Catalysts 13 00154 g010
Table 1. Recommended tests for photocatalytic materials according to the International Organization for Standardization.
Table 1. Recommended tests for photocatalytic materials according to the International Organization for Standardization.
Photocatalytic TestISO Reference
Anti-bacterial activityISO 27447:2019
Air purification
   Removal of NOISO 22197-1:2016
   Removal of acetaldehydeISO 22197-2:2019
   Removal of tolueneISO 22197-3:2019
Self-cleaning performanceISO 27448:2009
Water purification
   Generation of active oxygen speciesISO 10676:2010
   Degradation of methylene blueISO 10678:2010
   Dissolved oxygen consumption in a phenol solutionISO 19722:2017
Table 2. Selected examples of photocatalyst materials commonly synthesized for water treatment and green hydrogen production.
Table 2. Selected examples of photocatalyst materials commonly synthesized for water treatment and green hydrogen production.
Water Treatment
PhotocatalystModel MoleculeLight SourceReaction TimeOutcomesReference
TiO2TiO2 Degussa P-25Acridine orange and ethidium bromideSolar light-At pH = 10, it achieved the highest degradation rate[56]
Organic matter in petroleum wastewater3 hAt pH = 6.8, the reduction in chemical oxygen demand was of 76%[57]
Sol-gel TiO2Rhodamine BVisible light3.5 h60% of degradation rate[58]
TiO2/perliteFurfuralUV light2 h99.6% of degradation rate[59]
TiO2/corkCongo redUV light2 h85% of degradation rate[60]
TiO2/expanded polystyreneDrimaren red, indigo carmine, and methylene blueSolar light4 hAlmost 100% of degradation rate[61]
TiO2/low density polyethyleneMethylene blueUV light4 h35% of degradation rate[62]
Solar light30% of degradation rate
TiO2/LECAAmmoniaSolar light72 h96.5% of degradation rate[63]
Nano flower-like rutile TiO2Methylene blueSolar light3 h99% of degradation rate[64]
Multi-walled carbon nanotube/TiO22,4-dichlorophenolUV light2 h93% of degradation rate[65]
Solar light87% of degradation rate
Bi-doped TiO2 nanotubesRhodamine BSolar light4 h100% of degradation rate[66]
N-doped TiO2 nanotubes sintered at 350 °CMethyl orangeVisible light4 h96.3% of degradation rate[67]
B,N-codoped TiO2 supported on pearlstoneDiesel oilVisible light9 h48% of degradation rate[68]
Fe,N-codoped TiO2/fly ash cenospheresRhodamine BVisible light4 h89% of degradation rate[69]
C,N-codoped TiO2Rhodamine BVisible light1 h99% of degradation rate[70]
Chlorpyrifos2.5 h84% of degradation rate[71]
TiO2S-triazine herbicidesVisible light3 h 65% of degradation rate[72]
N-doped TiO2Crystal violet dyeUV light3 h~100% of degradation rate[73]
Ag-TiO2/corkRhodamine BSolar light4 h60% of degradation rate[74]
Fe3O4-TiO2Rhodamine BVisible light1 h100% of degradation rate[75]
Methylene blue75% of degradation rate
Congo red81% of degradation rate
Bi-doped TiO2Methyl orangeUV-visible light2 h94% of degradation rate[76]
TbxOy-TiO2PhenolVisible light3 hAt pH = 4, 100% was degraded[77]
Graphene oxide based TiO22-chlorophenolSolar light4 h80% of degradation rate[78]
Bisphenol AVisible light3 h67.6% of degradation rate[79]
ZnOZnORR141 azo dyeUV light4 h95% of degradation rate[80]
OfloxacinUV light5 h100% of degradation rate[81]
Solar light3.5 h82% of degradation rate[82]
Ag-ZnOOfloxacinSolar light3.5 h98% of degradation rate
Ag3PO4/ZnO nanowiresOfloxacinUV light0.5 h89% of degradation rate[83]
MnFe2O4-ZnOCongo redVisible light3.5 h54% of degradation rate[84]
ZnO@g-C3N4SulfamethoxazoleUV light0.5 h60% of degradation rate[85]
ZnO-SnO2Methylene blueUV light1 h95% of degradation rate[86]
ZnO/C3+-CeO2Methylene blueVisible light6 h97% of degradation rate[87]
ZnO-GrapheneMethyl orangeUV light4 h96% of degradation rate[88]
ZnO/PANIMethylene blueUV light3 h99% of degradation rate[89]
CeO2CeO2-CuOMethylene violetVisible light1.5 h93% of degradation rate[90]
CeO2/g-C3N4SulfamethoxazoleVisible light1 h99% of degradation rate[91]
CeO2/TiO2TetracyclineVisible light1.3 h100% of degradation rate[92]
Ag2O/CeO2EnrofloxacinVisible light2 h87% of degradation rate[93]
Ag/CeO2Rose bengal dyeVisible light3 h96% of degradation rate[94]
Pt/CeO2/ZnOPhenolVisible light1 h64% of degradation rate[95]
Co/CeO2Methylene blueVisible light6 h96% of degradation rate[96]
Cr/CeO2Methylene blueUV light1.6 h56% of degradation rate[97]
Fe/CeO2Methylene blueVisible light3 h97% of degradation rate[98]
Zr/CeO2Methylene blueVisible light2.5 h75% of degradation rate[99]
WO3ZnS-WO3-CoFe2O4Methylene blueVisible light3 h98% of degradation rate[100]
AgBr/WO3Rhodamine BVisible light2 h100% of degradation rate[101]
WO3/Bi2WO6Salicylic acid Visible light6 h75% of degradation rate[102]
BiVO4BiVO4OxytetracyclineVisible light2 h52% of degradation rate[103]
BiVO4/TiO2/RGOVisible light2 h68% of degradation rate[104]
AgI/BiVO4Visible light1 h80% of degradation rate[105]
Co-Pd/BiVO4PhenolVisible light3 h90% of degradation rate[106]
BiVO4/P-doped C3N4TetracyclineVisible light1 h97% of degradation rate[107]
RGO-TiO2/BiVO4Methylene blueVisible light2 h100% of degradation rate[108]
BiVO4/Bi2S3/MoS2Rhodamine BSolar light5 h97% of degradation rate[109]
Methylene blue93% of degradation rate
Malachite green94% of degradation rate
Bi2S3/BiVO4OxytetracyclineVisible light16 h67% of degradation rate[110]
Ag/Ag2S/BiVO4Visible light2.5 h100% of degradation rate[111]
Green hydrogen production through the water splitting reaction
PhotocatalystSacrificial agentLight sourceReaction timeOutcomesReference
TiO2Mesoporous N-TiO2 calcined at 250 °CMethanol (25% v/v)Visible light5 h31.5 μmol h−1 g−1[112]
Core shelled N-TiO2−xMethanol (20% v/v)Sunlight3.5 h15 mmol h−1 g−1[113]
Pt/N-TiO2Formic acid (10−3 mol L−1)Visible light6.5 h50.46 μmol h−1 g−1[114]
Ag/TiO2EthanolUV and visible light6 h1.34 μmol h−1 cm−2[115]
Ag/Ce-TiO2Ethanol (10% v/v)UV and visible light6 h1.47 mmol h−1 cm−2[116]
Fe,Ni-doped TiO2Ethanol (95% v/v)UV and visible light4 h24.3 mmol h−1 g−1[117]
Fe-doped and Cr-doped TiO2 thin films-Visible light6 h15.5 μmol h−1 cm−2[118]
Co-doped TiO2Glycerol (5% v/v)UV and solar light10 h11,021 μmol h−1 g−1[119]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Anaya-Rodríguez, F.; Durán-Álvarez, J.C.; Drisya, K.T.; Zanella, R. The Challenges of Integrating the Principles of Green Chemistry and Green Engineering to Heterogeneous Photocatalysis to Treat Water and Produce Green H2. Catalysts 2023, 13, 154. https://doi.org/10.3390/catal13010154

AMA Style

Anaya-Rodríguez F, Durán-Álvarez JC, Drisya KT, Zanella R. The Challenges of Integrating the Principles of Green Chemistry and Green Engineering to Heterogeneous Photocatalysis to Treat Water and Produce Green H2. Catalysts. 2023; 13(1):154. https://doi.org/10.3390/catal13010154

Chicago/Turabian Style

Anaya-Rodríguez, Fernanda, Juan C. Durán-Álvarez, K. T. Drisya, and Rodolfo Zanella. 2023. "The Challenges of Integrating the Principles of Green Chemistry and Green Engineering to Heterogeneous Photocatalysis to Treat Water and Produce Green H2" Catalysts 13, no. 1: 154. https://doi.org/10.3390/catal13010154

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop