Next Article in Journal
Advanced Electrical Measurements Technologies
Previous Article in Journal
Impact of Air Density Variation on a Simulated Earth-to-Air Heat Exchanger’s Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic Fuel Cells for Simultaneous Wastewater Treatment and Power Generation: Mechanisms, Challenges, and Future Prospects

1
Department of Chemistry, Amrit Campus, Tribhuvan University, Kathmandu 44618, Nepal
2
Department of Healthcare Management, Woosong University, Daejeon 34606, Korea
3
Carbon Composite Energy Nanomaterials Research Center, Woosuk University, Wanju, Chonbuk 55338, Korea
4
Woosuk Institute of Smart Convergence Life Care (WSCLC), Woosuk University, Wanju, Chonbuk 55338, Korea
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(9), 3216; https://doi.org/10.3390/en15093216
Submission received: 31 March 2022 / Revised: 25 April 2022 / Accepted: 25 April 2022 / Published: 28 April 2022

Abstract

:
Technological advancement is accompanied by excessive consumption of fossil fuels and affluent uses of chemical substances in many sectors, including transportation and manufacturing companies, and so on. Being an exhaustible resource, the excessive use of fossil fuels and of chemical substances may lead to a serious energy crisis in the long run, and it may additionally impose environmental pollution. Attempts have been made in the solution of such serious issues from every nook and corner. Nonetheless, no method has been found to be a panacea in waste water treatment and subsequent beneficiaries. One of the attempts in the solution to such issues is the application of photocatalytic technology, which could serve as a dual function in environmental remediation and clean energy production. A photocatalytic fuel cell is a tool developed for the recovery of energy from organic wastes. A rational cell construction needs the fabrication of photoelectrodes, the design of a photoanode and a photocathode chamber, in addition to an ion-transport membrane for pollution treatment and electricity generation. In this review, comprehensive fundamental assessments and recent developments in the design of photocatalytic fuel cells, their applications, future prospects, and challenges are covered.

1. Introduction

Large scale chemical synthesis, their derivatization, and commercial applications lag with rampant development, industrialization, and urbanization, which ultimately has consequences for the generation of varieties of complicated pollutants or synthetic compounds. Cumulative and continual disposal of such hazardous chemical substances into natural water bodies can pose an immense threat to human health and the environment [1,2,3]. The benefit of rampant availability of various synthetic chemical substances for a luxury life and the present world technology has created the burden of their appropriate disposal [4]. Traditionally, such solid and liquid effluents were being disposed into natural water sources like lakes, rivers, and oceans while gaseous effluents were disposed into the open sky as per the dictum ‘dilution is the solution of the pollution’ [5]. These days, the paradigm has shifted with the dawn of subsequent reflective problems and excessive and continual waste production. The present day’s talk is about the uses of non-hazardous materials with the proper disposal of waste materials in the vein of green chemistry. Therefore, this view wakes up to develop a new technique to dispose of waste materials with simultaneous energy generation. The constituents of waste materials could be heavy metal ions, organic compounds, and many other hybrid compounds [6,7,8]. The majority of the chemical waste obtained from dyes, textile, and chemical industries are organic compounds [9,10]. Photocatalysts can be used to degrade such organic materials. Hunge et al., developed an ultrasound treated graphene oxide/titanium oxide composite as a photocatalyst for the degradation of salicylic acid under sunlight illumination [11]. However, photocatalytic degradation of organic compounds is not the end story in the field of electrochemistry. Rather, a simultaneous action of photocatalytic degradation of organic compound and energy production is possible.
Organic matter can serve as an important source of chemically stored energy. In one facet, the presence of organic compounds in the air, water, or soil, poses a great environmental problem as waste material, and in another facet it serves as an important source for the recovery of energy from wastes. In this context “filling two needs with one deed” is the exploitation of chemical energy at the cost of organic waste degradation. Microbial fuel cell (MFC) can be a green approach for the utilization of waste organic materials for bioelectricity [12,13,14]. For decades, such bio-based electrochemical technologies have been studied and operated for the simultaneous decomposition of organic waste materials and electricity generation with the aid of the microbial community. However, many such pollutants are recalcitrant and difficult to degrade biologically [15]. Besides this, stringent working conditions, complexity in bacterial cultivation, poor bio-electrochemical activities, a long time, etc., have led to the search for other alternative methods for the generation of energy from waste organic materials [16,17]. In this run, photocatalysis function has been coupled with fuel cell operation—“the marriage of photocatalysis and fuel cells”—to develop alternative technology, such as the photocatalytic fuel cell (PFC) [18,19].
A photocatalyst is a semiconductor that increases the rate of reaction. Photocatalysis is a light-induced green chemical pathway in which especially organic compounds are oxidized into other products in the presence of light [20,21,22,23]. The photocatalytic oxidation process has been widely employed in the degradation of organic pollutants in air and water [24,25,26,27,28]. Light is cost-free, perpetual, clean, and convenient energy. An electrochemical cell devised to convert chemical energy stored in organic materials into electricity under the irradiation of sunlight is commonly termed a photoelectrochemical cell (PEC). PEC comprises three types of cells, viz; a photovoltaic cell, a photoelectrosynthetic cell, and a photocatalytic cell [29]. In this review article, an overview of photoelectrodes used in PEC and PFC is given. Subsequently, electrodes and their methods of preparation with major features of PFC are discussed along with their operation mechanism. Finally, the applications, future prospects, and challenges of PFC are covered.

2. Development of Single and Dual Photoelectrode PEC

Initially, hydrogen gas was produced by a water reduction process using PEC with a single photocatalyst. Electrodeposited cuprous oxide protected with nanolayers of aluminum-doped zinc oxide and titanium oxide Cu2O/(Al/ZnO-TiO2/Pt) was used as a promising photocathode for water reduction. In this case, the cathode was activated by the electrodeposition of platinum nanoparticles [30]. Later on, two electrodes PEC system (FTO/WO3 nanosheet as photoanode and Cu/Cu2O nanowire/NiOx as photocathode) were used for hydrogen production by water reduction [31]. The different band gap of Cu2O (2.0 eV) and WO3 (2.6 eV) permits an efficient and complementary light absorption for sunlight-driven water splitting phenomena. However, hydrogen production via the PEC system demands an external electrical energy supply.
A bit different from the PEC, a fuel cell has been devised for the direct conversion of chemically stored energy into electrical energy. The fuel cell is an energy conversion electrochemical device, where electricity is generated at the cost of an underlying chemical reaction [32]. The incorporation of the photocatalytic oxidation process in fuel cell operation gives birth to a photocatalytic fuel cell (PFC). Therefore, at its highest performance, a photocatalytic fuel cell is noted for its trios function of simultaneous organic pollutant degradation, electricity generation, and hydrogen production. Compared to those of solar cells, PFC uses organic wastes as fuel for the generation of electricity [33].
At the same time, a solar-driven fuel cell was used for the generation of electricity from wastewater. For this purpose, a photocatalytic fuel cell with a single photoelectrode (e.g., BiOCl/Ti as photoanode and Pt as photocathode) was used for the degradation of organic pollutants and simultaneous electricity generation. Later on, dual photoelectrode (WO3/W or TiO2/Ti as photoanode and Cu2O/Cu as photocathode) functioning under visible light illumination was applied for simultaneous wastewater treatment and electricity generation [34]. Based on this background, the development of a solar-driven dual photoelectrode PFC has come to the fore for simultaneous organic pollutant degradation and electricity generation. Under the irradiation of light, photogenerated holes are the majority for oxidation in n-type semiconductors and photogenerated electrons are the majority for the reduction in p-type semiconductors. In the formulation of PFC, the n-type semiconductor material is made anode which is expected to exhibit good photocatalytic properties, and the p-type semiconductor is made cathode where hydrogen production is expected. For p-type semiconductors as the photocathode, the electrode potential ECB Vs NHE is required to be more negative than the normal hydrogen electrode potential (EH+/H2). A schematic diagram of a single photoelectrode with Bi-TiO2 coupled with Pt cathode for RhB dye degradation is shown in the Figure 1 [35].

3. Electrodes Used in PFC

A typical photocatalytic fuel cell consists of two electrodes out of which one or both can be a photoelectrode. A photoanode comprises an n-type of semiconductor photocatalyst which consists of sufficient photoexcited electrons of the n-type of semiconductor [36]. It has a higher fermi energy level and functions as a strong oxidizing electrode for the degradation of organic materials. It supplies photogenerated electrons toward the cathode. The photoanode performs a dual function as a photocatalyst for photocatalysis and as an electrocatalyst for fuel cell activities. As the photoanode has a significant role in the generation of photocurrents, the nanoarchitectonics of anode materials seem to be of paramount importance. In the PFC system, the values of bandgap energy (Eg), the recombination rate of photoelectrons, and the feature of materials play a vital role in its performance. The fast induced electrons and their recombination with a hole in the photoanode can limit the performance of photoelectrochemical reactions. The bandgap energy of anode materials can be tuned by doping or by composite formation [37,38]. Commonly used photoanode in PFC includes TiO2, ZnO, WO3, α-Fe2O3, BaTiO3, BiVO4, etc. [39,40,41,42,43,44,45,46]. Among them, TiO2 is the most studied photoanode material applicable for PFC [47,48,49]. An increased interest in the use of TiO2 as a photocatalyst is preferable due to its stable physical and chemical features, low cost, facile methods of preparation, non-toxicity, profound stability, and excellent optoelectronic properties. Titanium oxide has been presented in various forms such as nanofilm, nanopowder, nanotubes, nanorods, etc. [50,51,52]. Furthermore, the use of nanoporous material avails the high surface area and increased the efficacy of photoelectrochemical performance. Due to the relatively higher bandgap of TiO2 (3.0–3.2 eV), it can absorb approximately 4% solar radiation, necessitating UV absorption for photoexcitation of electrons. On the contrary, the photogenerated electrons are more prone to combine with the hole in TiO2 resulting in reduced photocatalytic degradation [53]. Against these constraints, it is imperative to develop a visible light-responsive TiO2 for its application in photoelectrochemical fuel cell. To overcome the limitations of TiO2, various efforts have been put to facilitate visible light absorption and enhance the charge carrier separation. Some of such efforts includes surface modification [54], metal/metal oxide doping [55,56], non-metal doping [57], dye sensitization [58], etc.
Doping or composite formation with TiO2 brings about a change in the band gap energy and causes a shift in the UV absorption region, resulting in enhanced photooxidative properties [59,60,61]. For instance, the Eg of TiO2/BiVO4 composite reported 2.5 eV, while that for pure TiO2 and pure BiVO4 were 3.2 eV and 2.35 eV, respectively. In this work, the composite exhibited better photooxidative properties in the degradation of stearic acid and Rhodamine B compared to that from either component [62]. Silver nanoparticles incorporated TiO2, CdS-ZnS quantum dot sensitized TiO2 lead to the absorption band shifting to the visible region to harvest solar radiation [63,64,65]. Furthermore, the calcination temperature of TiO2 also affects the UV absorption properties and influences the performance of the photoelectrochemical cell. Gorska et al. used several calcination temperatures ranging from 350 °C to 750 °C to prepare TiO2 photocatalysts and they observed the best photocatalytic activity in the TiO2 prepared at 450 °C [66].
Commonly, platinum or platinum-based materials are used as cathode in photoelectrochemical fuel cells. A photocathode in PFC is made up of a p-type of semiconductor. Cuprous oxide (Cu2O), hematite (α-Fe2O3), reduced graphene oxide (RGO), ZnO/CuO composite, CuO/Co3O4, etc., have been reported to be used as a photocathode in PFC [67]. It has a lower Fermi energy level. At the cathode, electrocatalyst or photocatalyst induces an oxygen reduction reaction (ORR) in the presence of oxygen (aerobic condition) and water reduction reaction (WRR) or hydrogen evolution reaction (HER) in the absence of oxygen (anaerobic condition) [68,69,70]. As a consequence, either water or hydrogen is produced at the cathode depending upon the presence or absence of oxygen, respectively. Both electrodes are in contact with electrolytes in an electrochemical assembly. The photogenerated electrons are transported to the cathode via an external circuit due to the existing potential difference between the photoanode and the cathode, resulting in the generation of electricity [71]. As a net reaction, light energy is used to convert chemical energy into electrical energy directly in the fuel cell. In some cases, a reduction reaction may not be spontaneous at the photocathode due to the development of overpotential between the photocathode and the electrolyte. Such a condition necessitates the use of an electrocatalyst such that it would create a large interface between these two phases, and it facilitates the electron exchange with the electrolyte. A commonly used electrocatalyst for PFC cathode is a nanostructured platinum-coated carbon black in which the reaction depends upon the availability of oxygen. Water is produced under aerobic conditions and a hydrogen evolution reaction occurs under anaerobic conditions. However, by all means, the use of platinum is hindered due to its aggregation and higher cost tendency. Therefore, material scientists are searching for its alternatives. One of the most popularly used photocatalysts is Cu2O (bandgap of 2.0 eV), which is good for water reduction [67]. Besides this, cupric oxide (band gap of approximately 1.5 eV) can behave as an effective photocathode [72]. Titanium oxide and their various composite materials have been developed for the degradation of organic materials. Hunge et al., developed the titanium oxide@nanodiamond composite for the photocatalytic degradation of bisphenol A [73]. Graphene modified TiO2 have shown a high performance for organic pollutant removal due to the several benefits offered by graphene [25,74,75,76,77,78]. According to Tan et al. [79], besides the role of graphene, the type and the number of functional groups of the organic also determines the overall photocatalytic activities.
In a typical study, organic wastewater or electrolytes such as sodium sulfate, sodium chloride, etc. are put into the reaction chamber as electrolytes. Any organic matter such as dyes (methylene blue, methylene green, Congo red, methyl orange), drugs, phenols, etc., can act as fuel in PFC [80,81]. Besides these, photoelectrical treatment of landfill leachate, oil mill, textile, pharmaceutical, and tannery wastewater with simultaneous degradation organic matters and electricity generation can be carried out [82]. Table 1 shows some reported photoelectrodes and their methods of preparation with major features.

4. Mechanism of Photocatalytic Fuel Cell

The working of the photocatalytic fuel cell can well be described in terms of the band theory of solids. As an isolated atom is characterized by filled and vacant atomic orbitals, their assembly (lattice containing many atoms) constitutes the form of closely spaced filled and vacant orbitals, essentially forming a continuous band. The filled bonding orbitals (Highest occupied molecular orbital, HOMO) constitute the valence band (VB) and the vacant antibonding orbitals (lowest unoccupied molecular orbital, LUMO) forms the conduction band (CB) [90]. The energy difference (in electron volt) between the top of the valence band and the bottom of the conduction band is termed the band gap energy (Eg). It is the band gap energy that strongly influences the optical and the electrical properties of the electrode materials in an electrochemical setup [91,92]. When the band gap is very small, the conduction band and the valence band overlap such that electrons can move from the valence band to the conduction band easily, as that happened in a metallic conductor.
For larger values of Eg, the valence band is almost filled and the conduction band is almost vacant. Thermal excitation of electrons from the valence band to the conduction band makes a possibility of conduction in the materials. This constitutes a hole (electron deficit center) at the valence band where rearrangement of electrons is possible resulting in the mobility of holes. In case the band gap is greater than 1.5 eV, only a few charge carriers are available, which are ineffective for electrical conductivity. The different band gap regions are associated with the strong absorption edges of the electromagnetic spectrum. For instance, a band gap of approximately 1.2 eV corresponds to the near infrared region and a band gap of approximately 2.0 eV corresponds to the mid-visible region. A band energy of approximately 3.0 eV corresponds from the blue region to the ultraviolet region. A composite formation results in the band gap shifting. [93]. The addition of dopants (n-type or p-type) can introduce holes in the valence band and electrons in the conduction band. An n-type of dopant avails electrons, while a p-type of dopant pulls the resulting electrons into the formation of holes [94].
Tuning of electrochemical response in semiconductor electrodes can be made by altering the impurity levels, surface treatment, or surface adsorption [95]. The semiconductor–solution interface is very crucial in the conversion of light energy into electrical energy by photo-assisting the electron transfer between the semiconductor and the chemical species in a solution [96,97]. Besides this, the potential of donor/acceptor redox couples has also a crucial role in the photocatalysis process. The Fermi level of photoexcited electrons in the CB should be higher than the energy level of the acceptor otherwise, the reducible chemical species is unable to receive electrons at the cathode. In addition, the level of the holes (h+) in the VB of the semiconductor electrode should be at a lower position than that of the donor [83]. In the PFC couple designed by Chen et al., the WO3/W acts as a photoanode, and Cu2O/Cu acts as a photocathode where organic compounds can be decomposed (oxidized) by the holes of WO3/W photoanode only due to its higher oxidation power (approximately +3.1 VNHE) [83]. The WO3/W photoanode is an n-type semiconductor where the Fermi energy level is near the EVB, and the Cu2O/Cu acts as a p-type semiconductor where fermi energy lies near the ECB. The schematic diagram is shown in Figure 2.
Both the photoelectrodes can generate electron/hole pairs. The Fermi energy level of WO3/W photoanode is lower (more negative) than that of Cu2O/Cu as a cathode. Therefore, an interior bias is produced to cause the migration of photoelectrons of WO3/W photoanode to combine with the hole of Cu2O/Cu photocathode via an external circuit. The holes of WO3/W photoanode as well as electrons of Cu2O/Cu photocathode can oxidize the organic matter (R) with simultaneous water cleaning and energy recovery.
Reaction of hole of photoanode: nh+ + R → R’ + nH+
Reaction at photocathode: 2H+ + 2e → H2
When the energy of irradiated light exceeds the bandgap energy, the photoexcited electrons can be promoted from VB to CB, which facilitates the oxidation reaction at the anode and the reduction reaction at the cathode.
Photoanode + hυ → h+ + e
During this process, the photoexcited electron (e) is transferred to the CB of the semiconductor electrode, while the photoexcited hole (h+) remains with the VB. The photoexcited charges (electrons and holes) migrate on the surface of the semiconductor electrode, and they accomplish the underlying redox reaction.

4.1. Photoreduction Reaction

The photoexcited electrons migrate to the cathode and may react with dissolved oxygen to produce superoxide radicals (O2) or hydrogen peroxide (H2O2) or water (H2O) or hydroxyl radical (OH).
O2 + eO2 (strong oxidant)
The transportation of photoexcited electrons to the counter electrode via an external circuit generates electricity. The transport of electrons from electrodes to the electrolyte solution is governed by the Butler–Volmer equation [98].
i i 0 = exp   α A n F R T η     exp   α C n F R T η  
where i and i0 are instantaneous current density and exchange current density (in Am−2 unit), αA and αC are transfer coefficient (unitless) for anodic and cathodic current, respectively, n is the number of electrons transferred, F is Faraday’s constant, R is the universal gas constant, T is the absolute temperature and η is the over potential (deviation from equilibrium potential due to electrode polarization), respectively.
Furthermore, hydrogen ions are consumed in the presence of oxygen to give the hydrogen peroxide or water in an acidic or neutral medium. These are the cases of oxygen reduction reaction (ORR).
O2 + 2 H+ + 2 e → H2O2        E° = + 0.68 V (Oxygen reduction reaction)
O2 + 4 H+ + 4 e → 2 H2O        E° = + 1.23 V (Oxygen reduction reaction)
However, water reduction reaction (WRR) occurs at a basic medium.
1 2 O 2   + H 2 O + 2   e     2   OH   ( Water   reduction   reaction )
In the absence of oxygen at the cathode, the reduction of hydrogen ions takes place to produce hydrogen gas. This is often referred to as a hydrogen evolution reaction (HER) [70].
2 H+ + 2 e → H2 (proton reduction, at acidic pH, in absence of oxygen)

4.2. Photooxidation Reaction

The photoexcited hole (h+) in the VB may react with water molecules or hydroxyl ions to produce free hydroxyl radicals (OH). The H+ ions produced in the photooxidation process are brought to the cathode by a diffusion process in the electrolyte.
H2O + h+OH + H+
OH + h+OH (strong oxidant)

4.3. Decomposition of Organic Compound

Such photogenerated holes can act as an oxidant to decompose organic matter in wastewater. As a consequence, reactive chemical species such as superoxide oxygen radicals, hydroxyl radicals, etc., decompose the organic compound due to their strong oxidizing action. Besides these, some of the photoexcited charges may undergo recombination with the production of heat [99].
Organic compound + OH → CO2 + H2O + oxidized organic compound
Organic   compound + O 2 CO 2 + H 2 O + oxidized   organic   compound
In the photocatalytic fuel cells, organic pollutants act as fuel, which are brought to decompose under the irradiation of light at photoanode whereby electrons are photoexcited, which are driven by the potential difference between photoanode and photocathode. The efficacy of photocatalytic fuel cells is governed by many factors such as choice of visible-light photoelectrodes, cell design, duel photoelectrode setup, and inter-relation of these factors. Therefore, optimization in the formulation of effective semiconductors and rational design of photocatalytic fuel cells for organic waste degradation and electricity generation is a crucial aspect in the design of photocatalytic fuel cells.
The power density in PFC can be calculated by the relation,
Power density (P) = current density (J) × voltage (V)
Power density (P) can be expressed in mW cm−2, current density (J) can be expressed in mA cm−2, and voltage (V) can be expressed in mV [100].

5. Recent Development in the Design of PFC

In the early days, one of the most frequently used photoanodes was titanium oxide and the cathode was platinum. Later on, modification in electrode materials was carried out for a more efficient and versatile photoelectrode. These days, findings show that a dual photoelectrode could be more effective in the performance of PFC. In the development of a photoelectrode, binary or ternary materials are synthesized in search of better performance [101,102]. Besides these, many techniques are incorporated in the designing and the functioning of PFC. Great attention has been paid to the selection of electrode materials and their design in the formulation of PFC that is capable of performing multiple functions, including wastewater treatment, photoelectricity generation, hydrogen evolution, and carbon dioxide reduction [103]. Some PFCs are also capable of removing heavy metal ions such as Cr(VI) [104].
The photocatalytic activities of the PFC device can be increased by developing optofluidic technology, which is derived from microfluidic technology and optic fluidic technology. He et al., developed an optofluidic photocatalytic fuel cell composed of bismuth vanadate (BiVO4) as photoanode and cupric oxide (Cu2O) as photocathode for the generation of electricity from wastewater treatment [86]. A solar responsive photocatalytic fuel cell with CdS-ZnS quantum dot sensitized TiO2 as anode and Pt coated carbon as cathode was developed by He et al. and applied for simultaneous wastewater treatment and electricity generation [65]. Lui et al. developed a flow-photocatalytic fuel cell using Ag-TiO2 as a photoanode and Pt/C as a photocathode. When fed with real brewery wastewater, a continuous power generation of 1.85 W m−2 under simulated light was observed [88].
Chen et al. developed WO3/W photoanode and Cu2O/Cu photocathode for the simultaneous wastewater treatment and generation of electricity [67].
To enhance the efficacy of PFC, Li et al. have designed a dye self-photosensitization photo fuel cell (DSPFC), using bismuthoxychloride/titanium as photoanode. In this case, BiOCl is not activated directly in the visible region due to its wide band gap. Adsorption of some dyes such as rhodamine B and methylene green onto the surface of BiOCl can cause self-photosensitization. Dye adsorbed onto the BiOCl surface is excited by the visible light whereby electrons are injected from excited dye to the CB of BiOCl. Such excited electrons are scavenged by molecular oxygen to give oxygen ion radicals (O2• −). Herein, the conduction band of BiOCl (−1.1 eV) is more negative than that of O2/ O2• − (–0.046 eV) and hence the oxidation activity of O2• − is high which is sufficient to cause decomposition of the dye [71]. The selection and the formulation of the photoanode and the cathode have been done in such a way that the output of PFC in terms of open-circuit voltage (Voc), short circuit current (Jsc), maximum power density (Pmax), and photocurrent density have been greatly considered. Some of the PFCs with the corresponding photoanode, cathode/photocathode, the Voc, Jsc, Pmax, and PCD values are given in Table 2.

6. Applications of PFC

PFC has been exploited for many purposes. Some of the notable applications are as follows.
a. 
Wastewater treatment
Organic materials present in wastewater owe a significant amount of chemical energy, which can be transformed into electrical energy. Primarily, microbial fuel cells are being used to retrieve the electrical energy from such chemical storages. However, due to the biorefractory nature, MFC is not an appropriate device to treat persistent organic compounds such as azo dyes. In these contexts, PFC systems are getting more preference over MFC due to simple operational methods, efficacy of organic matter degradation, and direct electron transfer. The major constituents of textiles discharges consist of methyl oranges, methylene blue, and Congo red. The PFC can treat azo dye effectively.
b. 
Recovery of chemical energy from organic matter
Photoexcitation onto the photocatalyst is accompanied by the formation of electrons and holes. The holes are consumed in the oxidation of organic matter at the photoanode and the photoexcited electrons are transported to the photocathode. The quantitative generation of electricity depends upon various factors such as the nature of the electrolyte, concentration of electrolyte, efficacy of photoanode and photocathode, light intensity, pH, etc. The short-circuit current, open-circuit voltage, and power density change with the change in different parameters like concentration of electrolyte, pH, nature of organic compounds, etc. [99]. Li Kan et al., showed the uses of real textile wastewaters as a fuel for the generation of electricity and employed as a bias voltage (0.6–0.75 V) in a PFC reactor [109].
c. 
Hydrogen production
Wu et al., developed a solar-driven PFC with dual photoelectrodes using TiO2NRs/FTO as photoanode and C/Cu2O/Cu as photoanode for simultaneous wastewater treatment and hydrogen production. The operating mechanism of dual photoelectrode PFC is given in Figure 3. The PFC showed superior performance for the degradation of phenol and hydrogen production. In this case, the total hydrogen production rate was 86.8 μmol cm−2 [34].
d. 
Carbon dioxide reduction
These days, PFC has found its application in the reduction of carbon dioxide gas. Xie et al. have developed a coupled system of photocatalytic reduction and a photocatalytic fuel cell. In this system, carbon dioxide is first reduced to hydrocarbon by photocatalytic reduction, and the as-obtained compound is used as a fuel in PFC to generate electricity directly [108]. In the work of Yu et al., the photocatalytic fuel cell has been fueled with carbon dioxide saturated methanol-sulphuric acid solution. The design could hinder the recombination of electron-hole pairs and enhanced fuel cell performance in this engineering has been found [110]. As in the Figure 4, the CH3OH and H2SO4 mixed solution saturated with CO2 was supplied to the photoanode and the H2SO4 solution was supplied to the cathode.

7. Future Prospect and Challenges

Effective cell design and the architecture of photoanode has a crucial impact on the efficacy of photocatalytic fuel cell as it affects the generation of current by the transport of electron from photoanode to cathode. Furthermore, the conductive substrate-coated photocatalytic semiconductor can play a substantial role in the development of photocatalytic fuel cells with profound light-harvesting capacity. Conductive polymers such as polyaniline, polypyrrole, polythiophene, [111,112] etc., can be used in the development of composite materials either using with metallic nanoparticles or with nanofibers [113]. Surface functionalization can be performed to introduce some desired properties [114]. To enhance the surface area, nanostructured semiconductor conductive materials coated can be used with designed nanoarchitectonics. Such designs are required to be environmentally friendly and available at a low price. The generation of non-toxic byproducts in a PFC system could be supportive from the view of green chemistry. Some of the challenges in the synthesis of efficient PFC are their stability, purity, ease of synthesis, and cost factor. In some cases, the performance of PFC has been found to be reduced due to poor stability and photocorrosion, poor photoexcited charge separation, etc. For efficient performance, these methods and materials are required to exhibit high efficiency and good recycling stability. The development of PFC responding to both UV and visible light is a critical benefit in the successful formulation of PFC. Besides this, cost factor (using Pt as an electrode), toxicity concern (salt of cadmium), stability, and reusability are major issues to be considered in the development of PFC. Besides this, agglomeration is a major challenge in photocatalysis.

8. Conclusions and Recommendations

In real-world practice, the stability and the reusability of photoelectrodes is crucial for their sustainability and economic practicality. For the use of crystalline materials, XRD analysis of electrode materials before and after a certain duration of uses reveals the state of stability and reusability of electrodes. If the crystalline phase of the electrode materials remains almost constant even after several consecutive periods, the electrode is said to be stable, and it can be used successively. Further research work is necessary to increase the longevity of electrode materials. Uses of nanomaterial-based electrodes can avail a large surface area, inducing a higher rate of surface reaction. State-of-the-art nanotechnology and surface science can be applied to increase cell efficacy. Nanoarchitechtonics involves nano-creation and nano-organization with self-assembly, molecular manipulation, and chemical manipulation. Sustainable use of electrode material is necessary from the economic point of view.

Funding

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2020R1I1A1A01066994). This work was also supported by the Traditional Culture Convergence Program through the National Research Foundation of Korea (NRF) grant funded by the Ministry of Science, ICT and Future Planning (2018 M3C1B5052283).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

BPABisphenol A
CBConduction band
CODChemical oxygen demand
DSPFCDye self-photosensitization photo fuel cell
EgBand gap
FTOFluorine doped tin oxide
FFFill factor
HERHydrogen evolution reaction
HOMOHighest occupied molecular orbital
IscShort circuit current
JscShort circuit current density
(JV) maxMaximum power density
LUMOLowest unoccupied molecular orbital
LDHLayered double hydroxide
MBMethylene blue
MFCMicrobial fuel cell
MPDMaximum power density
OCVOpen circuit voltage
ORROxygen reduction reaction
PCDPhotocurrent density
PECPhotoelectrocatalytic cell
PECPhotoelectrochemical cells
PFCPhotocatalytic fuel cell
PLDPulsed laser deposition
RGOReduced graphene oxide
SCCShort circuit current
VBValence band
VocOpen circuit voltage
WRRWater reduction reaction

References

  1. Xu, P.; Xu, H.; Zheng, D. Simultaneous electricity generation and wastewater treatment in a photocatalytic fuel cell integrating electro-Fenton process. J. Power Sources 2019, 421, 156–161. [Google Scholar] [CrossRef]
  2. Pant, B.; Park, M.; Park, S.-J. Recent Advances in TiO2 Films Prepared by Sol-Gel Methods for Photocatalytic Degradation of Organic Pollutants and Antibacterial Activities. Coatings 2019, 9, 613. [Google Scholar] [CrossRef] [Green Version]
  3. Pant, B.; Barakat, N.A.M.; Pant, H.R.; Park, M.; Saud, P.S.; Kim, J.-W.; Kim, H.-Y. Synthesis and photocatalytic activities of CdS/TiO2 nanoparticles supported on carbon nanofibers for high efficient adsorption and simultaneous decomposition of organic dyes. J. Colloid Interface Sci. 2014, 434, 159–166. [Google Scholar] [CrossRef] [PubMed]
  4. Bhattarai, D.P.; Pant, B.; Acharya, J.; Park, M.; Ojha, G.P. Recent Progress in Metal–Organic Framework-Derived Nanostructures in the Removal of Volatile Organic Compounds. Molecules 2021, 26, 4948. [Google Scholar] [CrossRef]
  5. Hayworth, J.S.; Clement, T.P. BP’s Operation Deep Clean—Could Dilution be the Solution to Beach Pollution? Environ. Sci. Technol. 2011, 45, 4201–4202. [Google Scholar] [CrossRef]
  6. Marí-Guaita, J.; Bouich, A.; Marí, B. Shedding Light on Phase Stability and Surface Engineering of Formamidinium Lead Iodide (FaPbI3) Thin Films for Solar Cells. Eng. Proc. 2021, 12, 1. [Google Scholar]
  7. Bouich, A.; Marí-Guaita, J.; Bouich, A.; Pradas, I.G.; Marí, B. Towards Manufacture Stable Lead Perovskite APbI3 (A = Cs, MA, FA) Based Solar Cells with Low-Cost Techniques. Eng. Proc. 2021, 12, 81. [Google Scholar]
  8. Chang, X.; Fang, J.; Fan, Y.; Luo, T.; Su, H.; Zhang, Y.; Lu, J.; Tsetseris, L.; Anthopoulos, T.D.; Liu, S.; et al. Printable CsPbI3 Perovskite Solar Cells with PCE of 19% via an Additive Strategy. Adv. Mater. 2020, 32, 2001243. [Google Scholar] [CrossRef]
  9. Mani, S.; Bharagava, R.N. Textile industry wastewater: Environmental and health hazards and treatment approaches. In Recent Advances in Environmental Management; CRC Press: Boca Raton, FL, USA, 2018; pp. 47–69. [Google Scholar]
  10. Sharma, S.; Bhattacharya, A. Drinking water contamination and treatment techniques. Appl. Water Sci. 2017, 7, 1043–1067. [Google Scholar] [CrossRef] [Green Version]
  11. Hunge, Y.M.; Yadav, A.A.; Dhodamani, A.G.; Suzuki, N.; Terashima, C.; Fujishima, A.; Mathe, V.L. Enhanced photocatalytic performance of ultrasound treated GO/TiO2 composite for photocatalytic degradation of salicylic acid under sunlight illumination. Ultrason. Sonochem 2020, 61, 104849. [Google Scholar] [CrossRef]
  12. Chaturvedi, V.; Verma, P. Microbial fuel cell: A green approach for the utilization of waste for the generation of bioelectricity. Bioresour. Bioprocess. 2016, 3, 38. [Google Scholar] [CrossRef] [Green Version]
  13. Li, C.; Omine, K.; Sivasankar, V.; Sano, H.; Chicas, S.D. Development of low-cost solid phase microbial fuel cell using organic waste and recycling of materials after power generation: Characterization of carbon anode. Biomass Bioenergy 2021, 154, 106266. [Google Scholar] [CrossRef]
  14. Li, B.; Xu, D.; Feng, L.; Liu, Y.; Zhang, L. Advances and prospects on the aquatic plant coupled with sediment microbial fuel cell system. Environ. Pollut. 2022, 297, 118771. [Google Scholar] [CrossRef]
  15. Dhawle, R.; Mantzavinos, D.; Lianos, P. UV/H2O2 degradation of diclofenac in a photocatalytic fuel cell. Appl. Catal. B Environ. 2021, 299, 120706. [Google Scholar] [CrossRef]
  16. Logan, B.E.; Hamelers, B.; Rozendal, R.; Schröder, U.; Keller, J.; Freguia, S.; Aelterman, P.; Verstraete, W.; Rabaey, K. Microbial Fuel Cells:  Methodology and Technology. Environ. Sci. Technol. 2006, 40, 5181–5192. [Google Scholar] [CrossRef] [PubMed]
  17. Bruno, L.B.; Jothinathan, D.; Rajkumar, M. Microbial Fuel Cells: Fundamentals, Types, Significance and Limitations. In Microbial Fuel Cell Technology for Bioelectricity; Sivasankar, V., Mylsamy, P., Omine, K., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 23–48. [Google Scholar] [CrossRef]
  18. Lu, S.; Hummel, M.; Gu, Z.; Gu, Y.; Cen, Z.; Wei, L.; Zhou, Y.; Zhang, C.; Yang, C. Trash to treasure: A novel chemical route to synthesis of NiO/C for hydrogen production. Int. J. Hydrogen Energy 2019, 44, 16144–16153. [Google Scholar] [CrossRef]
  19. Selihin, N.M.; Tay, M.G. A review on future wastewater treatment technologies: Micro-nanobubbles, hybrid electro-Fenton processes, photocatalytic fuel cells, and microbial fuel cells. Water Sci. Technol. 2021, 85, 319–341. [Google Scholar] [CrossRef]
  20. Ameta, R.; Solanki, M.S.; Benjamin, S.; Ameta, S.C. Chapter 6—Photocatalysis. In Advanced Oxidation Processes for Waste Water Treatment; Ameta, S.C., Ameta, R., Eds.; Academic Press: Cambridge, MA, USA, 2018; pp. 135–175. [Google Scholar] [CrossRef]
  21. Wong, R.J.; Scott, J.; Low, G.K.C.; Feng, H.; Du, Y.; Hart, J.N.; Amal, R. Investigating the effect of UV light pre-treatment on the oxygen activation capacity of Au/TiO2. Catal. Sci. Technol. 2016, 6, 8188–8199. [Google Scholar] [CrossRef]
  22. Wong, R.J.; Scott, J.; Kappen, P.; Low, G.K.C.; Hart, J.N.; Amal, R. Enhancing bimetallic synergy with light: The effect of UV light pre-treatment on catalytic oxygen activation by bimetallic Au–Pt nanoparticles on a TiO2 support. Catal. Sci. Technol. 2017, 7, 4792–4805. [Google Scholar] [CrossRef]
  23. Wong, R.J.; Tsounis, C.; Scott, J.; Low, G.K.-C.; Amal, R. Promoting Catalytic Oxygen Activation by Localized Surface Plasmon Resonance: Effect of Visible Light Pre-treatment and Bimetallic Interactions. Chem. Cat Chem. 2018, 10, 287–295. [Google Scholar] [CrossRef]
  24. Pant, B.; Pant, H.R.; Barakat, N.A.M.; Park, M.; Jeon, K.; Choi, Y.; Kim, H.-Y. Carbon nanofibers decorated with binary semiconductor (TiO2/ZnO) nanocomposites for the effective removal of organic pollutants and the enhancement of antibacterial activities. Ceram. Int. 2013, 39, 7029–7035. [Google Scholar] [CrossRef]
  25. Pant, B.; Saud, P.S.; Park, M.; Park, S.-J.; Kim, H.-Y. General one-pot strategy to prepare Ag–TiO2 decorated reduced graphene oxide nanocomposites for chemical and biological disinfectant. J. Alloy. Compd. 2016, 671, 51–59. [Google Scholar] [CrossRef]
  26. He, F.; Jeon, W.; Choi, W. Photocatalytic air purification mimicking the self-cleaning process of the atmosphere. Nat. Commun. 2021, 12, 2528. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, D.; Cheng, Y.; Zhou, N.; Chen, P.; Wang, Y.; Li, K.; Huo, S.; Cheng, P.; Peng, P.; Zhang, R.; et al. Photocatalytic degradation of organic pollutants using TiO2-based photocatalysts: A review. J. Clean. Prod. 2020, 268, 121725. [Google Scholar] [CrossRef]
  28. Pant, B.; Park, M.; Kim, H.-Y.; Park, S.-J. Ag-ZnO photocatalyst anchored on carbon nanofibers: Synthesis, characterization, and photocatalytic activities. Synth. Met. 2016, 220, 533–537. [Google Scholar] [CrossRef]
  29. Bard, A.J. Photoelectrochemistry and heterogeneous photo-catalysis at semiconductors. J. Photochem. 1979, 10, 59–75. [Google Scholar] [CrossRef]
  30. Paracchino, A.; Laporte, V.; Sivula, K.; Grätzel, M.; Thimsen, E. Highly active oxide photocathode for photoelectrochemical water reduction. Nat. Mater. 2011, 10, 456–461. [Google Scholar] [CrossRef]
  31. Lin, C.-Y.; Lai, Y.-H.; Mersch, D.; Reisner, E. Cu2O|NiOx nanocomposite as an inexpensive photocathode in photoelectrochemical water splitting. Chem. Sci. 2012, 3, 3482–3487. [Google Scholar] [CrossRef]
  32. Larminie, J.; Dicks, A.; McDonald, M.S. Fuel Cell Systems Explained; J. Wiley: Chichester, UK, 2003; Volume 2. [Google Scholar]
  33. Lianos, P. Production of electricity and hydrogen by photocatalytic degradation of organic wastes in a photoelectrochemical cell: The concept of the Photofuelcell: A review of a re-emerging research field. J. Hazard. Mater. 2011, 185, 575–590. [Google Scholar] [CrossRef]
  34. Wu, Z.; Zhao, G.; Zhang, Y.; Liu, J.; Zhang, Y.-n.; Shi, H. A solar-driven photocatalytic fuel cell with dual photoelectrode for simultaneous wastewater treatment and hydrogen production. J. Mater. Chem. A 2015, 3, 3416–3424. [Google Scholar] [CrossRef]
  35. Liu, D.; Li, C.; Zhao, C.; Nie, E.; Wang, J.; Zhou, J.; Zhao, Q. Efficient Dye Contaminant Elimination and Simultaneously Electricity Production via a Bi-Doped TiO2 Photocatalytic Fuel Cell. Nanomaterials 2022, 12, 210. [Google Scholar] [CrossRef] [PubMed]
  36. Li, B.; He, Y.; Xiao, M.; Zhang, Y.; Wang, Z.; Qin, Z.; Chai, B.; Yan, J.; Li, J.; Li, J. A solar-light driven photocatalytic fuel cell for efficient electricity generation and organic wastewater degradation. Colloids Surf. A Physicochem. Eng. Asp. 2022, 642, 128205. [Google Scholar] [CrossRef]
  37. Pant, B.; Park, M.; Lee, J.H.; Kim, H.-Y.; Park, S.-J. Novel magnetically separable silver-iron oxide nanoparticles decorated graphitic carbon nitride nano-sheets: A multifunctional photocatalyst via one-step hydrothermal process. J. Colloid Interface Sci. 2017, 496, 343–352. [Google Scholar] [CrossRef] [PubMed]
  38. Wu, X.; Ng, Y.H.; Wen, X.; Chung, H.Y.; Wong, R.J.; Du, Y.; Dou, S.X.; Amal, R.; Scott, J. Construction of a Bi2MoO6:Bi2Mo3O12 heterojunction for efficient photocatalytic oxygen evolution. Chem. Eng. J. 2018, 353, 636–644. [Google Scholar] [CrossRef]
  39. Queiroz, B.D.; Fernandes, J.A.; Martins, C.A.; Wender, H. Photocatalytic fuel cells: From batch to microfluidics. J. Environ. Chem. Eng. 2022, 10, 107611. [Google Scholar] [CrossRef]
  40. Hu, C.; Kelm, D.; Schreiner, M.; Wollborn, T.; Mädler, L.; Teoh, W.Y. Designing Photoelectrodes for Photocatalytic Fuel Cells and Elucidating the Effects of Organic Substrates. Chem. Sus. Chem. 2015, 8, 4005–4015. [Google Scholar] [CrossRef]
  41. Imran, M.; Yousaf, A.B.; Farooq, M.; Kasak, P. Enhancement of visible light-driven hydrogen production over zinc cadmium sulfide nanoparticles anchored on BiVO4 nanorods. Int. J. Hydrogen Energy 2022, 47, 8327–8337. [Google Scholar] [CrossRef]
  42. 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]
  43. Mishra, P.; Saravanan, P.; Packirisamy, G.; Jang, M.; Wang, C. A subtle review on the challenges of photocatalytic fuel cell for sustainable power production. Int. J. Hydrogen Energy 2021, 46, 22877–22906. [Google Scholar] [CrossRef]
  44. Chung, H.Y.; Toe, C.Y.; Chen, W.; Wen, X.; Wong, R.J.; Amal, R.; Abdi, F.F.; Ng, Y.H. Manipulating the Fate of Charge Carriers with Tungsten Concentration: Enhancing Photoelectrochemical Water Oxidation of Bi2WO6. Small 2021, 17, 2102023. [Google Scholar] [CrossRef]
  45. Tan, H.L.; Tahini, H.A.; Wen, X.; Wong, R.J.; Tan, X.; Iwase, A.; Kudo, A.; Amal, R.; Smith, S.C.; Ng, Y.H. Interfacing BiVO4 with Reduced Graphene Oxide for Enhanced Photoactivity: A Tale of Facet Dependence of Electron Shuttling. Small 2016, 12, 5295–5302. [Google Scholar] [CrossRef] [PubMed]
  46. Tan, H.L.; Wen, X.; Amal, R.; Ng, Y.H. BiVO4 {010} and {110} Relative Exposure Extent: Governing Factor of Surface Charge Population and Photocatalytic Activity. J. Phys. Chem. Lett. 2016, 7, 1400–1405. [Google Scholar] [CrossRef] [PubMed]
  47. Mahmoud, M.; El-Kalliny, A.S.; Squadrito, G. Stacked titanium dioxide nanotubes photoanode facilitates unbiased hydrogen production in a solar-driven photoelectrochemical cell powered with a microbial fuel cell treating animal manure wastewater. Energy Convers. Manag. 2022, 254, 115225. [Google Scholar] [CrossRef]
  48. Kaneko, M.; Gokan, N.; Katakura, N.; Takei, Y.; Hoshino, M. Artificial photochemical nitrogen cycle to produce nitrogen and hydrogen from ammonia by platinized TiO2 and its application to a photofuel cell. Chem. Commun. 2005, 12, 1625–1627. [Google Scholar] [CrossRef]
  49. Liu, Y.; Li, J.; Zhou, B.; Chen, H.; Wang, Z.; Cai, W. A TiO2-nanotube-array-based photocatalytic fuel cell using refractory organic compounds as substrates for electricity generation. Chem. Commun. 2011, 47, 10314–10316. [Google Scholar] [CrossRef] [PubMed]
  50. Bhattarai, D.P.; Shrestha, S.; Shrestha, B.K.; Park, C.H.; Kim, C.S. A controlled surface geometry of polyaniline doped titania nanotubes biointerface for accelerating MC3T3-E1 cells growth in bone tissue engineering. Chem. Eng. J. 2018, 350, 57–68. [Google Scholar] [CrossRef]
  51. Rezk, A.I.; Bhattarai, D.P.; Park, J.; Park, C.H.; Kim, C.S. Polyaniline-coated titanium oxide nanoparticles and simvastatin-loaded poly (ε-caprolactone) composite nanofibers scaffold for bone tissue regeneration application. Colloids Surf. B Biointerfaces 2020, 192, 111007. [Google Scholar] [CrossRef]
  52. Leong, S.; Razmjou, A.; Wang, K.; Hapgood, K.; Zhang, X.; Wang, H. TiO2 based photocatalytic membranes: A review. J. Membr. Sci. 2014, 472, 167–184. [Google Scholar] [CrossRef]
  53. Li, M.; Liu, Y.; Dong, L.; Shen, C.; Li, F.; Huang, M.; Ma, C.; Yang, B.; An, X.; Sand, W. Recent advances on photocatalytic fuel cell for environmental applications—The marriage of photocatalysis and fuel cells. Sci. Total Environ. 2019, 668, 966–978. [Google Scholar] [CrossRef]
  54. Dubey, P.K.; Kumar, R.; Tiwari, R.S.; Srivastava, O.N.; Pandey, A.C.; Singh, P. Surface modification of aligned TiO2 nanotubes by Cu2O nanoparticles and their enhanced photo electrochemical properties and hydrogen generation application. Int. J. Hydrogen Energy 2018, 43, 6867–6878. [Google Scholar] [CrossRef]
  55. Tao, S.; Wang, F.; Zhang, J.; Shi, J.; Guo, W.; Lu, J. Visible-Light-Responsive TiO2/NiFe Mixed Metal Oxide-Carbon Photocatalytic Fuel Cell with Synchronous Hydrogen Peroxide Production. Eur. J. Inorg. Chem. 2021, 2021, 1230–1239. [Google Scholar] [CrossRef]
  56. Marami, M.B.; Farahmandjou, M.; Khoshnevisan, B. Sol–Gel Synthesis of Fe-Doped TiO2 Nanocrystals. J. Electron. Mater. 2018, 47, 3741–3748. [Google Scholar] [CrossRef]
  57. Yu, H.; Zhang, M.; Wang, Y.; Lv, J.; Liu, Y.; He, G.; Sun, Z. Low-temperature strategy for vapor phase hydrothermal synthesis of C\N\S-doped TiO2 nanorod arrays with enhanced photoelectrochemical and photocatalytic activity. J. Ind. Eng. Chem. 2021, 98, 130–139. [Google Scholar] [CrossRef]
  58. Thor, S.-H.; Ho, L.-N.; Ong, S.-A.; Abidin, C.Z.A.; Heah, C.-Y.; Nordin, N.; Ong, Y.-P.; Yap, K.-L. Advanced oxidation treatment of amaranth dye synchronized with electricity generation using carbon-based cathodes in a sustainable photocatalytic fuel cell integrated electro-fenton system. J. Environ. Chem. Eng. 2021, 9, 106439. [Google Scholar] [CrossRef]
  59. Pant, B.; Prasad Ojha, G.; Acharya, J.; Park, M. Ag3PO4-TiO2-Carbon nanofiber Composite: An efficient Visible-light photocatalyst obtained from eelectrospinning and hydrothermal methods. Sep. Purif. Technol. 2021, 276, 119400. [Google Scholar] [CrossRef]
  60. Pant, B.; Park, M.; Park, S.-J. Hydrothermal synthesis of Ag2CO3-TiO2 loaded reduced graphene oxide nanocomposites with highly efficient photocatalytic activity. Chem. Eng. Commun. 2020, 207, 688–695. [Google Scholar] [CrossRef]
  61. Khlyustova, A.; Sirotkin, N.; Kusova, T.; Kraev, A.; Titov, V.; Agafonov, A. Doped TiO2: The effect of doping elements on photocatalytic activity. Mater. Adv. 2020, 1, 1193–1201. [Google Scholar] [CrossRef]
  62. Monfort, O.; Roch, T.; Gregor, M.; Satrapinskyy, L.; Raptis, D.; Lianos, P.; Plesch, G. Photooxidative properties of various BiVO4/TiO2 layered composite films and study of their photocatalytic mechanism in pollutant degradation. J. Environ. Chem. Eng. 2017, 5, 5143–5149. [Google Scholar] [CrossRef]
  63. Sui, M.; Dong, Y.; Wang, Z.; Wang, F.; You, H. A biocathode-driven photocatalytic fuel cell using an Ag-doped TiO2/Ti mesh photoanode for electricity generation and pollutant degradation. J. Photochem. Photobiol. A Chem. 2017, 348, 238–245. [Google Scholar] [CrossRef]
  64. Wang, B.; Zhang, H.; Lu, X.-Y.; Xuan, J.; Leung, M.K.H. Solar photocatalytic fuel cell using CdS–TiO2 photoanode and air-breathing cathode for wastewater treatment and simultaneous electricity production. Chem. Eng. J. 2014, 253, 174–182. [Google Scholar] [CrossRef]
  65. He, X.; Chen, M.; Chen, R.; Zhu, X.; Liao, Q.; Ye, D.; Zhang, B.; Zhang, W.; Yu, Y. A solar responsive photocatalytic fuel cell with the membrane electrode assembly design for simultaneous wastewater treatment and electricity generation. J. Hazard. Mater. 2018, 358, 346–354. [Google Scholar] [CrossRef]
  66. Górska, P.; Zaleska, A.; Kowalska, E.; Klimczuk, T.; Sobczak, J.W.; Skwarek, E.; Janusz, W.; Hupka, J. TiO2 photoactivity in vis and UV light: The influence of calcination temperature and surface properties. Appl. Catal. B Environ. 2008, 84, 440–447. [Google Scholar] [CrossRef]
  67. Chen, Q.; Li, J.; Li, X.; Huang, K.; Zhou, B.; Cai, W.; Shangguan, W. Visible-Light Responsive Photocatalytic Fuel Cell Based on WO3/W Photoanode and Cu2O/Cu Photocathode for Simultaneous Wastewater Treatment and Electricity Generation. Environ. Sci. Technol. 2012, 46, 11451–11458. [Google Scholar] [CrossRef]
  68. Zeng, Q.; Chang, S.; Wang, M.; Li, M.; Deng, Q.; Xiong, Z.; Zhou, B.; Liu, Y. Highly-active, metal-free, carbon-based ORR cathode for efficient organics removal and electricity generation in a PFC system. Chin. Chem. Lett. 2021, 32, 2212–2216. [Google Scholar] [CrossRef]
  69. Sfaelou, S.; Lianos, P. Photoactivated Fuel Cells (PhotoFuelCells). An alternative source of renewable energy with environmental benefits. AIMS Mater. Sci. 2016, 3, 270–288. [Google Scholar] [CrossRef]
  70. Khalik, W.F.; Ho, L.-N.; Ong, S.-A.; Voon, C.-H.; Wong, Y.-S.; Yusuf, S.Y.; Yusoff, N.; Lee, S.-L. Reactive Black 5 as electron donor and/or electron acceptor in dual chamber of solar photocatalytic fuel cell. Chemosphere 2018, 202, 467–475. [Google Scholar] [CrossRef] [PubMed]
  71. Li, K.; Xu, Y.; He, Y.; Yang, C.; Wang, Y.; Jia, J. Photocatalytic fuel cell (PFC) and dye self-photosensitization photocatalytic fuel cell (DSPFC) with BiOCl/Ti photoanode under UV and visible light irradiation. Environ. Sci. Technol. 2013, 47, 3490–3497. [Google Scholar] [CrossRef]
  72. Masudy-Panah, S.; Siavash Moakhar, R.; Chua, C.S.; Kushwaha, A.; Dalapati, G.K. Stable and Efficient CuO Based Photocathode through Oxygen-Rich Composition and Au–Pd Nanostructure Incorporation for Solar-Hydrogen Production. ACS Appl. Mater. Interfaces 2017, 9, 27596–27606. [Google Scholar] [CrossRef]
  73. Hunge, Y.M.; Yadav, A.A.; Khan, S.; Takagi, K.; Suzuki, N.; Teshima, K.; Terashima, C.; Fujishima, A. Photocatalytic degradation of bisphenol A using titanium dioxide@nanodiamond composites under UV light illumination. J. Colloid Interface Sci. 2021, 582, 1058–1066. [Google Scholar] [CrossRef]
  74. Pant, H.R.; Pant, B.; Pokharel, P.; Kim, H.J.; Tijing, L.D.; Park, C.H.; Lee, D.S.; Kim, H.Y.; Kim, C.S. Photocatalytic TiO2–RGO/nylon-6 spider-wave-like nano-nets via electrospinning and hydrothermal treatment. J. Membr. Sci. 2013, 429, 225–234. [Google Scholar] [CrossRef]
  75. Trapalis, A.; Todorova, N.; Giannakopoulou, T.; Boukos, N.; Speliotis, T.; Dimotikali, D.; Yu, J. TiO2/graphene composite photocatalysts for NOx removal: A comparison of surfactant-stabilized graphene and reduced graphene oxide. Appl. Catal. B Environ. 2016, 180, 637–647. [Google Scholar] [CrossRef]
  76. Tang, B.; Chen, H.; Peng, H.; Wang, Z.; Huang, W. Graphene Modified TiO2 Composite Photocatalysts: Mechanism, Progress and Perspective. Nanomaterials 2018, 8, 105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Pant, H.R.; Adhikari, S.P.; Pant, B.; Joshi, M.K.; Kim, H.J.; Park, C.H.; Kim, C.S. Immobilization of TiO2 nanofibers on reduced graphene sheets: Novel strategy in electrospinning. J. Colloid Interface Sci. 2015, 457, 174–179. [Google Scholar] [CrossRef]
  78. Pant, B.; Park, M.; Park, S.-J.; Kim, H.-Y. One-pot synthesis of CdS sensitized TiO2 decorated reduced graphene oxide nanosheets for the hydrolysis of ammonia-borane and the effective removal of organic pollutant from water. Ceram. Int. 2016, 42, 15247–15252. [Google Scholar] [CrossRef]
  79. Tan, H.L.; Denny, F.; Hermawan, M.; Wong, R.J.; Amal, R.; Ng, Y.H. Reduced graphene oxide is not a universal promoter for photocatalytic activities of TiO2. J. Mater. 2017, 3, 51–57. [Google Scholar] [CrossRef]
  80. Bai, J.; Wang, R.; Li, Y.; Tang, Y.; Zeng, Q.; Xia, L.; Li, X.; Li, J.; Li, C.; Zhou, B. A solar light driven dual photoelectrode photocatalytic fuel cell (PFC) for simultaneous wastewater treatment and electricity generation. J. Hazard. Mater. 2016, 311, 51–62. [Google Scholar] [CrossRef]
  81. Kee, M.W.; Soo, J.W.; Lam, S.M.; Sin, J.C.; Mohamed, A.R. Evaluation of photocatalytic fuel cell (PFC) for electricity production and simultaneous degradation of methyl green in synthetic and real greywater effluents. J. Environ. Manag. 2018, 228, 383–392. [Google Scholar] [CrossRef]
  82. Divyapriya, G.; Singh, S.; Martínez-Huitle, C.A.; Scaria, J.; Karim, A.V.; Nidheesh, P.V. Treatment of real wastewater by photoelectrochemical methods: An overview. Chemosphere 2021, 276, 130188. [Google Scholar] [CrossRef]
  83. Matarrese, R.; Mascia, M.; Vacca, A.; Mais, L.; Usai, E.M.; Ghidelli, M.; Mascaretti, L.; Bricchi, B.R.; Russo, V.; Casari, C.S.; et al. Integrated Au/TiO2 Nanostructured Photoanodes for Photoelectrochemical Organics Degradation. Catalysts 2019, 9, 340. [Google Scholar] [CrossRef] [Green Version]
  84. Ye, Y.; Bruning, H.; Li, X.; Yntema, D.; Rijnaarts, H.H.M. Significant enhancement of micropollutant photocatalytic degradation using a TiO2 nanotube array photoanode based photocatalytic fuel cell. Chem. Eng. J. 2018, 354, 553–562. [Google Scholar] [CrossRef]
  85. Lee, S.-L.; Ho, L.-N.; Ong, S.-A.; Wong, Y.-S.; Voon, C.-H.; Khalik, W.F.; Yusoff, N.A.; Nordin, N. A highly efficient immobilized ZnO/Zn photoanode for degradation of azo dye Reactive Green 19 in a photocatalytic fuel cell. Chemosphere 2017, 166, 118–125. [Google Scholar] [CrossRef] [PubMed]
  86. He, Y.; Zhang, Y.; Li, L.; Shen, W.; Li, J. An efficient optofluidic photocatalytic fuel cell with dual-photoelectrode for electricity generation from wastewater treatment. J. Solid State Chem. 2021, 293, 121780. [Google Scholar] [CrossRef]
  87. Hunge, Y.M.; Mahadik, M.A.; Bulakhe, R.N.; Yadav, S.P.; Shim, J.J.; Moholkar, A.V.; Bhosale, C.H. Oxidative degradation of benzoic acid using spray deposited WO3/TiO2 thin films. J. Mater. Sci. Mater. Electron. 2017, 28, 17976–17984. [Google Scholar] [CrossRef]
  88. Lui, G.; Jiang, G.; Fowler, M.; Yu, A.; Chen, Z. A high performance wastewater-fed flow-photocatalytic fuel cell. J. Power Sources 2019, 425, 69–75. [Google Scholar] [CrossRef]
  89. Nahyoon, N.A.; Liu, L.; Rabe, K.; Thebo, K.H.; Yuan, L.; Sun, J.; Yang, F. Significant photocatalytic degradation and electricity generation in the photocatalytic fuel cell (PFC) using novel anodic nanocomposite of Fe, graphene oxide, and titanium phosphate. Electrochim. Acta 2018, 271, 41–48. [Google Scholar] [CrossRef]
  90. Scharber, M.C.; Sariciftci, N.S. Low Band Gap Conjugated Semiconducting Polymers. Adv. Mater. Technol. 2021, 6, 2000857. [Google Scholar] [CrossRef]
  91. Sitt, A.; Hadar, I.; Banin, U. Band-gap engineering, optoelectronic properties and applications of colloidal heterostructured semiconductor nanorods. Nano Today 2013, 8, 494–513. [Google Scholar] [CrossRef]
  92. Sinha, B.; Goswami, T.; Paul, S.; Misra, A. The impact of surface structure and band gap on the optoelectronic properties of Cu2O nanoclusters of varying size and symmetry. RSC Adv. 2014, 4, 5092–5104. [Google Scholar] [CrossRef]
  93. Vojkovic, S.; Fernandez, J.; Elgueta, S.; Vega, F.E.; Rojas, S.D.; Wheatley, R.A.; Seifert, B.; Wallentowitz, S.; Cabrera, A.L. Band gap determination in multi-band-gap CuFeO2 delafossite epitaxial thin film by photoconductivity. SN Appl. Sci. 2019, 1, 1322. [Google Scholar] [CrossRef] [Green Version]
  94. Persson, C.; Lindefelt, U.; Sernelius, B.E. Band gap narrowing in n-type and p-type 3C-, 2H-, 4H-, 6H-SiC, and Si. J. Appl. Phys. 1999, 86, 4419–4427. [Google Scholar] [CrossRef]
  95. Haram, S.K. 9—Semiconductor Electrodes. In Handbook of Electrochemistry; Zoski, C.G., Ed.; Elsevier: Amsterdam, The Netherlands, 2007; pp. 329–389. [Google Scholar] [CrossRef]
  96. Kohl, P.A.; Bard, A.J. Semiconductor Electrodes: XVII. Electrochemical Behavior of n-and p-Type Electrodes in Acetonitrile Solutions. J. Electrochem. Soc. 1979, 126, 598. [Google Scholar] [CrossRef]
  97. Pant, B.; Ojha, G.P.; Kim, H.-Y.; Park, M.; Park, S.-J. Fly-ash-incorporated electrospun zinc oxide nanofibers: Potential material for environmental remediation. Environ. Pollut. 2019, 245, 163–172. [Google Scholar] [CrossRef] [PubMed]
  98. Dickinson, E.J.; Wain, A.J. The Butler-Volmer equation in electrochemical theory: Origins, value, and practical application. J. Electroanal. Chem. 2020, 872, 114145. [Google Scholar] [CrossRef]
  99. He, Y.; Chen, K.; Leung, M.K.; Zhang, Y.; Li, L.; Li, G.; Xuan, J.; Li, J. Photocatalytic fuel cell–A review. Chem. Eng. J. 2022, 428, 131074. [Google Scholar] [CrossRef]
  100. Vasseghian, Y.; Khataee, A.; Dragoi, E.-N.; Moradi, M.; Nabavifard, S.; Conti, G.O.; Khaneghah, A.M. Pollutants degradation and power generation by photocatalytic fuel cells: A comprehensive review. Arabian J. Chem. 2020, 13, 8458–8480. [Google Scholar] [CrossRef]
  101. Ammar, S.H.; Shafi, R.F.; Ali, A.D. A novel airlift photocatalytic fuel cell (APFC) with immobilized CdS coated zerovalent iron (Fe@ CdS) and g-C3N4 photocatalysts film as photoanode for power generation and organics degradation. Colloids Surf. A Physicochem. Eng. Asp. 2020, 602, 125164. [Google Scholar] [CrossRef]
  102. Li, L.; Chen, R.; Zhu, X.; Liao, Q.; Ye, D.; Zhang, B.; He, X.; Jiao, L.; Feng, H.; Zhang, W. A ternary hybrid CdS/SiO2/TiO2 photoanode with enhanced photoelectrochemical activity. Renew. Energy 2018, 127, 524–530. [Google Scholar] [CrossRef]
  103. Ješić, D.; Lašič Jurković, D.; Pohar, A.; Suhadolnik, L.; Likozar, B. Engineering photocatalytic and photoelectrocatalytic CO2 reduction reactions: Mechanisms, intrinsic kinetics, mass transfer resistances, reactors and multi-scale modelling simulations. Chem. Eng. J. 2021, 407, 126799. [Google Scholar] [CrossRef]
  104. Liu, X.-H.; Xing, Z.-H.; Chen, Q.-Y.; Wang, Y.-H. Multi-functional photocatalytic fuel cell for simultaneous removal of organic pollutant and chromium (VI) accompanied with electricity production. Chemosphere 2019, 237, 124457. [Google Scholar] [CrossRef]
  105. Yao, H.; Xu, Y.; Zhong, D.; Zeng, Y.; Zhong, N. Efficient rhodamine B degradation and stable electricity generation performance of visible-light photocatalytic fuel cell with g-C3N4/WO3/TiO2/Ti photoanode. Ionics 2021, 27, 4875–4884. [Google Scholar] [CrossRef]
  106. He, Y.; Yuan, R.; Yan, J.; Li, J. A highly efficient NiFe-layer double hydroxide/TiO2 heterojunction photoanode-based high-performance bifunctional photocatalytic fuel cell. Mater. Today Commun. 2021, 26, 102177. [Google Scholar] [CrossRef]
  107. Yong, Z.-J.; Lam, S.-M.; Sin, J.-C.; RahmanMohamed, A. Feasibility study of municipal wastewater removal synchronized with electricity generation via solar-driven photocatalytic fuel cell with Bi2WO6/ZnO nanorods array photoanode. IOP Conf. Ser. Earth Environ. Sci. 2021, 945, 012004. [Google Scholar]
  108. Xie, F.; Chen, R.; Zhu, X.; Liao, Q.; Ye, D.; Zhang, B.; Yu, Y.; Li, J. CO2 utilization: Direct power generation by a coupled system that integrates photocatalytic reduction of CO2 with photocatalytic fuel cell. J. CO2 Util. 2019, 32, 31–36. [Google Scholar] [CrossRef]
  109. Li, K.; Zhang, H.; Tang, T.; Xu, Y.; Ying, D.; Wang, Y.; Jia, J. Optimization and application of TiO2/Ti–Pt photo fuel cell (PFC) to effectively generate electricity and degrade organic pollutants simultaneously. Water Res. 2014, 62, 1–10. [Google Scholar] [CrossRef] [PubMed]
  110. Yu, Y.; Xie, F.; Chen, R.; Zhu, X.; Liao, Q.; Ye, D.; Li, J.; Song, S. New insights into the role of CO2 in a photocatalytic fuel cell. J. Power Sources 2021, 487, 229438. [Google Scholar] [CrossRef]
  111. Bhattarai, D.P.; Hwang, T.I.; Kim, J.I.; Lee, J.H.; Chun, S.; Kim, B.-S.; Park, C.H.; Kim, C.S. Synthesis of polypyrrole nanorods via sacrificial removal of aluminum oxide nanopore template: A study on cell viability, electrical stimulation and neuronal differentiation of PC12 cells. Mater. Sci. Eng. C 2020, 107, 110325. [Google Scholar] [CrossRef]
  112. Bhattarai, D.P.; Kim, B.S. NIR-triggered hyperthermal effect of polythiophene nanoparticles synthesized by surfactant-free oxidative polymerization method on colorectal carcinoma cells. Cells 2020, 9, 2122. [Google Scholar] [CrossRef]
  113. Tiwari, A.P.; Pandeya, S.; Bhattarai, D.P.; Joshi, M.K. Biomimetic Mineralization of Electrospun PCL-Based Composite Nanofibrous Scaffold for Hard Tissue Engineering. In Nanoscale Engineering of Biomaterials: Properties and Applications; Springer: Berlin/Heidelberg, Germany, 2022; pp. 683–704. [Google Scholar]
  114. Bhattarai, D.P.; Pokharel, P.; Xiao, D. Surface functionalization of polymers. In Reactive and Functional Polymers Volume Four; Springer: Berlin/Heidelberg, Germany, 2020; pp. 5–34. [Google Scholar]
Figure 1. Schematic diagram showing the structure of PFC connecting with the electrochemical workstation [35].
Figure 1. Schematic diagram showing the structure of PFC connecting with the electrochemical workstation [35].
Energies 15 03216 g001
Figure 2. Energy level of diagram of the PFC cell (WO3/W−Cu2O/Cu) for organic compounds degradation and electricity generation (Vp, photovoltage). Adapted with permission from Ref. [67]. Copyright Copyright 2012, American Chemical Society.
Figure 2. Energy level of diagram of the PFC cell (WO3/W−Cu2O/Cu) for organic compounds degradation and electricity generation (Vp, photovoltage). Adapted with permission from Ref. [67]. Copyright Copyright 2012, American Chemical Society.
Energies 15 03216 g002
Figure 3. A working diagram of PFC system. Adopted with the permission from Ref. [34]. Copyright The Royal Society of Chemistry 2014.
Figure 3. A working diagram of PFC system. Adopted with the permission from Ref. [34]. Copyright The Royal Society of Chemistry 2014.
Energies 15 03216 g003
Figure 4. Schematic illustration of a PFC operating with the CO2 saturated methanol-H2SO4 solution. Reprinted with the permission form ref. [110]. Copyright 2021 Elsevier B.V.
Figure 4. Schematic illustration of a PFC operating with the CO2 saturated methanol-H2SO4 solution. Reprinted with the permission form ref. [110]. Copyright 2021 Elsevier B.V.
Energies 15 03216 g004
Table 1. Electrodes and their methods of preparation with major features of PFC.
Table 1. Electrodes and their methods of preparation with major features of PFC.
Photoanode
Materials
Cathode/PhotocathodeMethods of PreparationSpecial FeaturesPerformanceRef.
TiO2 NRs/FTOC/Cu2O NW As/CuHydrothermal/anodizationPhenol degradation and simultaneous hydrogen production TOC removal rate 84.2% and H2 production rate 86.8 μmol cm−2[34]
Au/TiO2PtPulsed laser deposition (PLD)Bisphenol A (BPA) oxidation, water splitting, hydrogen productionPhotocurrent density = 58 μA/cm2[83]
TiO2 nanotube arrayCuElectrochemical anodizationMicropollutant removal from water62% MCPA removal [84]
ZnO/ZnPt/CHeat attachment methodDegradation of reactive green 19100% removal of RG19, Jsc = 0.0427 mAcm−2, and
(JV)max = 0.0102 mW cm−2
[85]
BiVO4 bismuth vandateCu2O-Good stability,
Electricity generation and waste water treatment.
Voc = 0.463 V
and Jsc = 0.113 mA cm−2
[86]
CdS-ZnS-TiO2Pt black on carbon blackSuccessive ionic layer adsorption and reactionOrganic compound degradation and electricity generation.JVmax = 1.01 mW cm−2 at a current density of 1.4 mA cm−2. FF = 0.45[65]
WO3/WFe@Fe2O3/carbon felt electro-fentonHydrothermal processWaste water treatment and electricity generation. Jsc = 0.59 mA cm−2, JVmax = 0.34 mW cm−2[1]
WO3/TiO2-Spray pyrolysis methodPhotoelectrolytic degradation of benzoic acidBenzoic acid degradation rate of 46.56%,
Voc = 0.485 V
Isc = 0.575 mA
[87]
Ag-TiO2Pt/C on carbon clothSolvothermal processCoulombic efficiency of 9.4% JVmax = 1.85 Wm−2
COD removal = 14.8%
[88]
MoS2/TiO2MoS2/Ni foamHydrothermal processHigher light harvesting, electron transport efficiency, Rhodamine B (20 ppm, pH =7) degradation efficiency of 69.16%. Voc = 0.7 V
Isc = 1.04 mA cm−2
JVmax = 0.114 mW cm−2.
[36]
WO3/WCu2O/CuHydrothermal processWaste water treatment and electricity generationDegradation rate of phenol, Rhodamine B and Congo red were 58%, 63%, 74%, respectively.[67]
BiOCl/TiPtWet chemical synthesis and dip coating methodRhodamine B and salicylic acid degradation and electricity generationJsc = 0.0116, Voc = 0.655 V, ff = 0.39 for Salicylic acid under UV irradiation.[71]
Bi doped TiO2PtSol-gel methodRhodamine B degradation and photocurrent generation91.2% degradation of Rhodamine B[35]
NiFe layered double oxide/TiO2Carbon black co-precipitation/calcinationHydrogen peroxide productionVoc = 0.78 V
Jsc = 1093 μA cm−2.
Pmax = 169 μW cm−2.
[55]
BiVO4/TiO2 NTsZnO/CuO NWsHydrothermal/liquid phase depositionDegradations of methyl orange, Congo red and methylene blue are 76%, 83%, and 90%, respectively.JVmax = 0.116 mW cm−2.[80]
Fe-graphene oxide-titanium phosphateZnIn2S4Hummer’s modified method, wet chemical methodDegradation of Rhodamine B90% removal capacity for Rhodamine B at pH 1.[89]
Table 2. Some electrochemical parameters of MFC.
Table 2. Some electrochemical parameters of MFC.
SNPhotoanodeCathode/
Photocathode
OCV
(Voc)
SCC
(Jsc) mA cm−2
MPD
(Pmax) μW cm−2
PCD mA cm−2Degradation Profile Ref
1g-C3N4/WO3/TiO2/TiPt-0.20 19.35 2.27 at 1.1 V vs. SCERhB degradation by 87.7%[105]
2NiFe-layered double hydroxide/TiO2Cu2O0.6900.622206 1.85 at 1.6 V vs RHEMB degradation (10 ppm) 94.6%[106]
3Bi2WO6/ZnO nanorodsPt563 mv37.103.30 [107]
4ZnO/CPt/C1000 mV0.00692.34-Degradation of Azo dye reactive black 5 (13.6% at anode and 8.7% at cathode)[70]
5BiVO4/TiO2 NTsZnO/CuO NWs0.53 V0.430.116 mW cm−2-Degradation of Methyl orange (76%), Congo red (90%), and Methylene blue (83%)[80]
6TiO2NRs/FTOC/Cu2O/Cu0.41 V0.50--Phenol removal (84.2%)[34]
7ZnO/ZnPt/C1050 mV0.03270.0076 mW cm−2-Degradation of reactive green 19[85]
8TiO2/FTOTiO20.98 V0.1750.110 mW cm−2-CO2 utilization[108]
OCV = Open circuit voltage (Voc), SCC = short circuit current (Jsc), MPD = maximum power density (Pmax), PCD = Photocurrent density.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Oli, H.B.; Kim, A.A.; Park, M.; Bhattarai, D.P.; Pant, B. Photocatalytic Fuel Cells for Simultaneous Wastewater Treatment and Power Generation: Mechanisms, Challenges, and Future Prospects. Energies 2022, 15, 3216. https://doi.org/10.3390/en15093216

AMA Style

Oli HB, Kim AA, Park M, Bhattarai DP, Pant B. Photocatalytic Fuel Cells for Simultaneous Wastewater Treatment and Power Generation: Mechanisms, Challenges, and Future Prospects. Energies. 2022; 15(9):3216. https://doi.org/10.3390/en15093216

Chicago/Turabian Style

Oli, Hari Bhakta, Allison A. Kim, Mira Park, Deval Prasad Bhattarai, and Bishweshwar Pant. 2022. "Photocatalytic Fuel Cells for Simultaneous Wastewater Treatment and Power Generation: Mechanisms, Challenges, and Future Prospects" Energies 15, no. 9: 3216. https://doi.org/10.3390/en15093216

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