Next Article in Journal
Stability of Quantum-Dot Light Emitting Diodes with Alkali Metal Carbonates Blending in Mg Doped ZnO Electron Transport Layer
Previous Article in Journal
Biodegradable Hydrogels Loaded with Magnetically Responsive Microspheres as 2D and 3D Scaffolds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photons to Formate: A Review on Photocatalytic Reduction of CO2 to Formic Acid

Department of Chemistry, New Mexico Institute of Mining and Technology, Socorro, NM 87801, USA
*
Author to whom correspondence should be addressed.
Current Address: School of Chemistry and Materials Science, Rochester Institute of Technology, 84 Lomb Memorial Drive, Rochester, New York, NY 14623, USA.
Nanomaterials 2020, 10(12), 2422; https://doi.org/10.3390/nano10122422
Submission received: 26 September 2020 / Revised: 24 November 2020 / Accepted: 26 November 2020 / Published: 4 December 2020
(This article belongs to the Special Issue Nanocomposites for Photocatalytic CO2 and Bicarbonate Reduction)

Abstract

:
Rising levels of atmospheric carbon dioxide due to the burning and depletion of fossil fuels is continuously raising environmental concerns about global warming and the future of our energy supply. Renewable energy, especially better utilization of solar energy, is a promising method for CO2 conversion and chemical storage. Research in the solar fuels area is focused on designing novel catalysts and developing new conversion pathways. In this review, we focus on the photocatalytic reduction of CO2 primarily in its neutral pH species of carbonate to formate. The first two-electron photoproduct of carbon dioxide, a case for formate (or formic acid) is made in this review based on its value as; an important chemical feedstock, a hydrogen storage material, an intermediate to methanol, a high-octane fuel and broad application in fuel cells. This review focuses specifically on the following photocatalysts: semiconductors, phthalocyanines as photosensitizers and membrane devices and metal-organic frameworks.

1. Introduction

The sun provides 100,000 TW of energy in one hour that is enough to power the earth for one year, [1] but the energy needs to be storable and transportable (i.e., chemical fuel) because renewable energy sources generally provide an intermittent supply of energy [2]. In principle, all terrestrial energy sources, such as fossil fuels and uranium, can be considered as a solar fuel; it just depends on the time frame. For the purpose of this review, when solar energy is converted into chemical fuel on a diurnal basis the term solar fuels applies. As a mitigation strategy, C1 sources such as CO2 or its pH neutral dissolved version, bicarbonate become concentrated energy sources for long-term storage capacity via energy input from solar irradiation [3]. In 2019, fossil fuels still supply 84% of the world’s energy [4]. In addition to limitations in the availability of fossil fuels, the emission of greenhouse gases, especially carbon dioxide, upon the combustion of fossil fuels is a major contributor to rising levels of anthropogenic carbon dioxide in the atmosphere. With the limitations in availability of fossil fuels, new sources of energy that provide a large-scale, sustainable energy supply must be developed. Solar energy is the most abundant renewable energy source available that can meet our future energy demands [5]. One potential approach towards generating renewable fuels is to use solar energy to directly reduce atmospheric or locally produced CO2 to liquid fuels. This approach, referred to as “chemical carbon mitigation” can lead to methanol as an end product, a useful solar fuel. The concept of “the methanol economy,” championed by Chemistry Nobel laureate George Olah, highlights methanol as a renewable and readily transportable fuel and an alternative to the hydrogen economy. A key component to several C1 conversion methods in the methanol economy is the solar-driven conversion of carbon dioxide to formic acid and ultimately to methanol, a renewable and regenerative C1 fuel [6,7,8].
The photocatalytic reduction of CO2 is a multi-electron transfer process that can lead to the formation of many different products depending on the reaction pathway and the number of electrons transferred. Two-electron products include carbon monoxide, formic acid and oxalate. Methanol is the six-electron product and methane is the result of eight electron transfers [9,10,11,12,13]. The reduction half-reaction utilizes the photogenerated electrons, leaving behind the photogenerated holes. Ideally, the oxidation half-reaction would lead to the oxidation of water generating hydrogen or oxygen but very few systems can accomplish simultaneous oxidation and reduction reactions [9]. Therefore, research has focused on developing a sacrificial agent that can efficiently scavenge the photogenerated holes.
Photocatalysis follows three key steps (Figure 1): (1) Photon absorption by semiconductor, (2) charge separation, (3) surface reactions. Semiconductors are attractive materials for photocatalysis because their band gap energies are similar to that of UV or visible light (Figure 2) [9]. Upon absorption of a photon with energy greater or equal to the band gap of the semiconductor, an electron is excited from the valence band to the conduction band, leaving behind a hole. The photo-generated charge carriers then migrate to the semiconductor surface and transfer to the adsorbed molecules, which initiates the subsequent oxidation or reduction reactions. The formation of charge carriers is followed by several de-excitation pathways (Figure 1). Transfer of the charge carrier, either electron or hole, to the acceptor molecules leads to its oxidation or reduction (pathway 3 and 4). The carriers can recombine with the opposite charge carrier trapped at the surface (pathway 5) or recombination can occur in the bulk of the semiconductor (pathway 6). Three major drawbacks of photocatalysis include (1) narrow region of optical absorption, (2) rapid electron/hole recombination and (3) lack of understanding of photocatalysis mechanisms. This review aims to address these challenges.
This review is structured as follows: Section 2 and Section 3 discusses the fundamentals of photocatalytic CO2 reduction and explains why our work is focused on photocatalytic bicarbonate reduction to formate. Section 4 divides the catalysts into semiconductors, phthalocyanine-semiconductors and metal-organic frameworks. Lastly, Section 5 discusses the outlook of photochemical CO2 reduction to formate and Section 6 summarizes the review with concluding remarks.

2. Fundamentals of Photocatalytic CO2 Reduction

Thermodynamics of CO2 Reduction

Photocatalytic reduction of CO2 has a positive ΔGº value, meaning the process is non-spontaneous and requires significant energy input from the incident photons. As mentioned previously, methanol production from CO2 reduction requires six protons and six electrons. Recently, there are studies in which homogeneous sensitizers coupled with metal co-catalysts have been shown to produce only CO and formic acid [14]. This observation of the incomplete six-electron reduction to methanol is governed by thermodynamics. Methanol production in an aqueous medium at pH 7 occurs at a slightly more positive potential (−0.39 V vs. NHE) than formate (−0.58 V vs. NHE) (Table 1). In addition, reduction of CO2 to formate only requires two electrons whereas reduction to methanol requires six.
Since carbon is in its highest oxidized state in CO2, its reduction can lead to a wide variety of products (CO, CH4, CH3OH, HCOOH and many other products). CO2 is a thermodynamically stable and chemically inert molecule with the C=O bond possessing a high dissociation energy of ~750 kJ/mol and C-C bond of ~336 kJ/mol, indicating a significant amount of energy input [16,17]. Moreover, the breaking of the C=O bonds and the bending of the molecular from linear to bent has been shown to require a large input of energy [18]. In addition, Equation (9) has been shown to be the necessary step to activate CO2, which has a very negative potential; therefore, high overpotential is needed to convert a CO2 molecule to products [9,13].
CO2 + e → CO2•−   E° = −1.90 V vs. NHE
We chose to highlight bicarbonate reduction in this study because of its reduction efficiency over dissolved CO2 gas due to its limited solubility of CO2 in water (0.033 M at 298 K and 1 atm) [19]. As several reports require the use of aqueous conditions, bubbled CO2 poses a concentration disadvantage. In a previous study by our group [20], we questioned the standard assumption that CO2 was the only species undergoing reduction even though numerous publications were using a bicarbonate buffer. This paper was among the first to conduct a control experiment with only bicarbonate buffer and no bubbled CO2 [21,22]. From this control experiment, we discovered that bubbling CO2 did not result in significantly higher formate production. Consequently, we decided to focus our efforts on bicarbonate reduction due to the following major issues with CO2. First, a considerable practical problem with atmospheric capture is that the concentration of CO2 in the air is still fairly low at 100’s of ppm [23]. While several concentrated CO2 sources exist at fossil fuel burning plants or in some cases oil drilling activities, it is very hard to supply a reactor with enough CO2 if it is being drawn from the atmosphere. Our experimental conditions parallel ocean bicarbonate concentration of 0.033 M [19] and with the continual acidification of the ocean due to dissolved carbon dioxide, it is relevant to capture bicarbonate and convert it to a value-added product.

3. The Case for Photons to Formate

In comparison to hydrogen, fuels based on carbon hold the advantage when it comes to denser volumetric energy along with transportation and storage within the current industrialized infrastructure. Converting CO2 into value-added chemicals could contribute to negative emission technologies (NETs) by carbon sequestration into acetic acid, ethylene and formic acid instead of expensive geological storage [11]. Of these commodity chemicals, formate production and its growing chemical application has increased significantly over the past decade [24]. Furthermore, in relation to higher Cn homologs, CO2 reduction to specialty chemicals in the range of C1-C3 such as formic acid production (0.2 MtC/yr) or propanol (0.1 MtC/yr) offers a brighter outlook for near term economic sustainability [25,26,27]. By comparison, carbon mitigation at the GtC/yr scale, which is ultimately impactful toward global climate change, will necessitate chemical conversion of greenhouse gases to be competitive with the production of fossil fuels such as coal (4 GtC/yr) and natural gas (1.4 GtC/yr) [28,29].

Techno-Economic Analysis

Even though CO2 reduction is a widely studied field, much of the research is focused on the fundamental aspects of catalyst design and product selectivity. However, for this process to be commercialized, many requirements need to be met such as (1) continuous operation, (2) high product selectivity and throughput and (3) long-term stable operation [30]. In addition, it is essential that the energy used in the CO2 conversion process does not generate additional CO2. Recent studies on the economic feasibility of a commercial CO2 reduction reactor show that formic acid is one of the only products that has potential to be made on a large-scale basis. Jouny et al. calculated the end-of-life net present value (NPV) of a CO2 electrolyzer for the production of 100 tons/day of various CO2 reduction products and found that only carbon monoxide and formic acid were the only economically viable products with NPVs of $13.5 million and $39.4 million, respectively [25]. Spurgeon and Kumar analyzed electrochemical conversion of flue gas CO2 to liquid products and found that electrosynthesized formic acid from CO2 was analyzed to be near market prices as a bulk chemical, even though it is not promising as an economic fuel. The major issue was the high capital cost for CO2 electrolyzers since high Faradaic efficiency, high current densities and reduced electrolyzer cost is necessary for economic viability [31]. Rumayor et al. carried out a techno-economic analysis of electrochemical reduction of CO2 to formic acid and found that it is not yet profitable and competitive under current market conditions due to the high production costs [32].

4. Photocatalytic CO2 Reduction to Formic Acid

4.1. Semiconductors

Halmann and Inoue were the first to reduce CO2 with semiconductor photocatalysts. Within a year of each other, both published their works in Nature reported the photo-reduction of CO2 in aqueous medium. In 1978, Halmann constructed a photoelectrode consisting of p-type gallium phosphide that produced formic acid after 18 h of irradiation [33]. A year later, Inoue reported the reduction of CO2 using several photosensitive semiconductor powders (TiO2, ZnO, CdS, GaP, SiC and WO3) suspended in water [34]. Inoue’s work was one of the first studies to hypothesize that for photocatalysis to be efficient, photo-excited electrons in the more negative conduction band have greater ability to reduce carbon dioxide. Even though this work produced low quantum yields (0.0005%), it was an important early study that inspired many other researchers. Following these two papers, both authors continued to make novel discoveries in this field. Halmann, in 1982, demonstrated enhanced formate production could be achieved by using oxides of rare earth dopants (Eu2O3, Sm2O3, Nd2O3, CeO2) to dope large band gap semiconductors (BaTiO3, LiNbO3) [35]. The optical properties of the catalysts were studied by diffuse reflectance spectroscopy, which uncovered the fact that rare earth dopants allowed the extension of optical absorption in the visible region of approximately 600 nm. A year later, Halmann reported the photoelectrochemical reduction of aqueous CO2 using a single crystal p-gallium phosphide and p-gallium arsenide as photocathodes in neutral or weakly acidic medium [36]. The main product was determined to be formate (71 µmol) but formaldehyde and methanol were also produced. The influence of pH on the behavior of the electrodes was studied and a strong photo-effect was seen for all pH values. The next year, Halmann once again studied TiO2 in the photo-reduction of CO2, this time doped with RuO2 [37]. They found that by doping anatase TiO2 with RuO2, formate was the predominant product and efficiency reached 0.04%. Later in 1995, Inoue reported on the photoreduction of CO2 using chalcogenide semiconductor microcrystals. This study synthesized cadmium-loaded ZnS microcrystals in a sodium bicarbonate buffer and bubbled CO2, which produced the highest quantum efficiency of 32.5%, twice as large as that obtained on bare ZnS microcrystals. However, the quantum efficiency was determined using 280 nm irradiation, which would explain the high efficiency by using monochromatic light. 2-propanol was used as a hole scavenger and it was found that Cd loading was an important factor in formate production. It was also reported that with increasing mole fraction of Cd, formate production actually decreased and that the optimal Cd loading is 0.5–0.67%. Other metals such as Pb, Ni, Ag and Cu were used but none produced as high efficiency as Cd-ZnS [38].
Since the work by Inoue, ZnS has become an attractive semiconductor photocatalyst because of its non-toxic and earth abundant attributes and its effectiveness in the reduction of CO2. Kanemoto was able to achieve sufficient formate production of 75.1 µmol/h but this experiment was performed at 313 nm irradiation [39]. These findings also confirmed 2-propanol to be the most efficient electron donor amid sodium phosphate, sodium sulfate and triethanolamine. Johne and Kisch loaded ZnS onto large surface area SiO2 particles to produce formate under UV irradiation. Later, 2,5-dihydrofuran was used as a reducing agent, which assisted in the production of 7000 µmol of formate [40].
Our group recently explored the photoreduction of bicarbonate to formate using wurtzite and sphalerite ZnS particles [20]. ZnS was studied in depth under the parameters of size, crystal lattice, surface area and band gap on the productivity of formate. Photochemical experiments were performed under air mass coefficient (AM 1.5 and 0) solar simulator conditions. Formate production was negligible under AM 1.5 conditions but significantly increased under AM 0 conditions due to the inclusion of shorter wavelength photons. This work was one of the first to question the standard assumption that CO2 is the only species undergoing reduction. Compared to earlier works reporting bubbled CO2 in the presence of bicarbonate, this work suggests that bicarbonate was the predominant species undergoing reduction. As the C-H bond undergoes oxidation in 2-propanol, glycerol with three hydroxylic positive hole scavengers was hypothesized as an improved semiconductor hole-scavenger and green chemistry solvent glycerol greatly improved the efficiency of the reaction. Continuing with ZnS, our group studied copper (I) oxide (Cu2O) in the photoreduction of bicarbonate to formate. This work examined micron- and nano-particulate Cu2O and synthesized a Ag/Cu2O nanocomposite [41]. Due to plasmon-induced electron transfer from silver to Cu2O, an enhanced formate production was found with the Ag/Cu2O nanocomposite photocatalyst at 5.5% apparent quantum efficiency (Figure 3). In addition, an Au/TiO2 and Ag/TiO2 nanocomposite photocatalyst could reach apparent quantum efficiencies of 2.4% and 7.8%, respectively [42,43]. Beierle et al. designed a titanium nitride-titanium dioxide nanocomposite for bicarbonate reduction to formate. This work demonstrated that under solar illumination, the TiN-TiO2 nanocomposite has higher photocatalytic activity than bare TiN or TiO2, this nanocomposite can also greatly enhance formate production and remain stable for 8 h [44].
Baran et al. synthesized nanocrystalline zinc sulfide surface-modified with ruthenium(0) for the photocatalytic CO2 reduction to formate, carbon monoxide and methane [45]. Formation of acetone was also observed due to the oxidation of the hole scavenger isopropanol. Surface-modified ZnS was more efficient than pure ZnS due to (1) the lowering of the CO2 activation energy, (2) collection of electrons at Ru particles resulted in a better charge separation and (3) better CO2 desorption at the surface of modified ZnS. Both polar and non-polar solvents were tested and it was reported that in a polar solvent (water), formate is the major product and water was a better H-transfer agent than isopropanol. Kuwabata et al. employed ZnS microcrystallites for photocatalytic CO2 reduction to methanol and formate by using methanol dehydrogenase and (MDH) and pyrroloquinoline quinone (PQQ) as an electron mediator for MDH and 2-propanol as a positive hole scavenger [46]. Acetone production resulting from the oxidation of 2-propanol was also studied.
Irvine et al. used CdS, ZnO, SiC, BaTiO3 and SrTiO3 to photochemically reduce aqueous CO2 [47]. Product analysis showed the formation of formate, formaldehyde and methanol; the highest efficiency was obtained with ZnO, CdS and SiC. Continuing on Irvine’s work, Eggins et al. used CdS or ZnS colloids for the photo-reduction of CO2 to produce formate and other dimeric and tetrameric products such as oxalate, glyoxylate, glycolate and tartrate. Other semiconductors studied included ZnO, SiC, BaTiO3 and SrTiO3 [48]. This work discusses the role of several hole acceptor (or electron donor) compounds that assist in reacting with photogenerated holes. Different electron donors tested include iron ferracyanide, hydroquinone, 2-propanol, RuO2 and hypophosphite. This paper highlights that semiconductors with the most negative conduction band potential results in the best quantum yields because they have a stronger reducing potential. This work concluded that ZnS resulted in the highest quantum yield among all the semiconductors tested. Henglein et al. reported the efficient photoreduction of CO2 to formic acid using SiO2-stabilized ZnS and 2-propanol as electron donor. By employing this system 0.80% AQE was achieved [49].
Kaneco and coworkers published three studies, all employing TiO2 in a liquid CO2 medium [50,51,52]. In the first effort, the effect of temperature, pressure, illumination time and surface area were examined. It was found that the formate yield was neither affected by temperature nor pressure. Formate production increased sharply with illumination time up until 5 h and progressed slowly until 30 h. A similar trend was observed with the surface area of the catalyst, formate increased linearly with increasing surface area of TiO2 up to 0.4 m2/g. Electron spin resonance elucidated the reaction mechanism of the reaction, which identified two radical species, photogenerated Ti3+ and CO2•− radical anion which resulted from the first electron transfer to CO2. In their follow-up work, Kaneco and coworkers continued to investigate the photoreduction of high-pressure CO2 using TiO2 powders with 2-propanol as a positive hole scavenger [51]. Methane was found to be the main reduction product; formate was also detected but only under extremely high pressures of CO2. In 1999, Kaneco and coworkers continued to study TiO2 powders in supercritical fluid CO2 by the investigation on the effects of temperature, pressure, irradiation time, surface area and aqueous solution for protonation. This study determined that formate production increased with an acidic pH medium because H+ are advantageous for desorption of reaction intermediates from the surface of TiO2 [52]. These two works were the first to point out that photo-excited TiO2 results in (Ti3+-O)* species (Figure 4) [51,52].
Liu et al. studied photocatalytic CO2 reduction on CdS particles with and without surface modification in various solvents [53]. Formate and carbon monoxide were the two major products, as well as acetone from the oxidation of 2-propanol. The type of solvent and surface modification played a major role in the selectivity of the products. The reduction of CO2 in acetonitrile produced carbon monoxide and formate, while in dichloromethane produced carbon monoxide only. Kisch reported CdS supported on silica and zinc sulfide to photocatalyze the reduction of bicarbonate to formate [54]. For CdS supported on zinc sulfide, it was found that the optimal CdS loading was 5%, which increased formate production 40-fold and 16-fold as compared to bare CdS and ZnS. Formate production was found to be dependent on bicarbonate concentration, when bicarbonate concentration increased from 33 to 60 mM formate production was enhanced by 10%. Formate production decreased by 20% when the pH was lowered to 5.3 but at a pH of 11.2, formate production increased by 25%. Zhang et al. used manganese sulfide (MnS) to reduce bicarbonate to formate at neutral pH [55]. Both ZnS and MnS have their conduction bands at highly reducing positions at −1.04 and −1.19 V versus NHE. This work results in quantum efficiency of 4.2%, however, the wavelengths utilized fell within the UV region of the solar spectrum, between 200 and 400 nm.
Xia et al. synthesized a multi-walled carbon nanotube (MWCNT) supported TiO2 and found that the addition of MWCNT greatly enhanced the photocatalytic activity of TiO2 (Figure 5) [56]. As with other doped catalysts, it was found that there is an optimal amount of MWCNT before it begins to decrease the photocatalytic activity of TiO2 due to the fact that excessive MWCNT can shield the TiO2 from absorbing UV light. The optimal amount of MWCNT prevents TiO2 from aggregating and can delay electron/hole recombination. Qin et al. made a bifunctionalized TiO2 film containing a dye-sensitized zone and a catalysis zone designed for visible light photocatalytic CO2 reduction (Figure 6) [57]. Electrons transferred from the dye to the conduction band of TiO2 resulted in highly efficient conversion of CO2 to formate, formaldehyde and methanol. After the electron injection, the dye is oxidized to dye+, which can be regenerated by accepting electrons from I from the electrolyte. Ulagappan et al. studied the photoreduction of gaseous CO2 in Ti-silicate molecular sieves using methanol as an electron donor [58]. The main result of the study is the detection of formate and it was proposed that upon excitation of the metal center, transient Ti+III reduces CO2 while methanol is oxidized.
In the 2000s, doping, co-catalyst and extending optical absorption to the visible region became a popular area for research. Sato et al. coupled a p-type semiconductor (N-doped Ta2O5) with a ruthenium complex as an electrocatalyst to produce formate from CO2 [59]. This was the first report of electron transfer from semiconductor in the excited state to a metal complex in the ground state. Triethanolamine (TEOA) was used as an electron donor as well as a proton source and the measured quantum yield was 1.9% at 405 nm. A year later, Sato et al. studied the photoelectrochemical reduction of CO2 over p-type InP/Ru complex polymer hybrid photocatalyst [60]. This reduction, using H2O as an electron donor and proton source was achieved by using a two-step excitation process called a Z-scheme (Figure 7). Highly selective CO2 photo-reduction was achieved by Arai and coworkers by combining Cu2ZnSnS4 (CZTS) with a metal complex electrocatalyst [61]. The main product was identified to be formate, which demonstrates the selectivity and efficiency of the hybrid semiconductor-metal-complex-electrocatalyst catalyst. Suzuki et al. continued on that idea of p-type photoactive N-doped Ti2O5 with a ruthenium complex, which can efficiently reduce CO2 to formate using visible light [62].
Co-catalysts became a popular method to enhance semiconductor efficiency. Iizuka et al. used Ag as a co-catalyst for ALa4Ti4O15 (where A = Ca, Sr and Ba) to reduce CO2 to formate [63]. No sacrificial agent was used because water was consumed as a reducing agent (electron donor). It was found that the optimum Ag loading was 2 wt% and that BaLa4Ti4O15 was the most active photocatalyst. Palladium is another noble metal that is often used as a co-catalyst. Raja et al. synthesized two photocatalysts-anodized titanium oxide nanotubes (T-NT) and bismuth titanate (BTO) decorated with palladium nanoparticles [64]. Pd-BTO showed a 2-fold increase in formic acid (110 µmol/h/g) yield when compared to bare BTO or T-NT, which is attributed to better charge separation in the hybrid Pd-BTO and the higher visible light harvesting ability of BTO because of its lower bandgap.
Mendoza et al. employed Co3O4 powders for CO2 photo-reduction under visible light irradiation [65]. Co3O4 is a p-type semiconductor with band gap of 2.0 eV and has a conduction band edge that is more negative than the potentials of the CO2 reduction reactions. In this study, formate was produced without any hole scavenger nor photo-sensitizer. Qin et al. synthesized a bismuth yttrium oxide (BiYO3) photocatalyst for CO2 reduction to formate under visible light irradiation [66]. Due to its large surface area and smaller band gap, BiYO3 produced fewer hydroxyl radicals to give a higher formate yield.
Most recently, copper oxide and copper oxide-derivatives have been gaining attention as a highly active and earth abundant photocatalyst. Ali and coworkers designed a Z-scheme heterostructure composed of reduced titania and Cu2O capable of reducing CO2 to CH4. Due to the Z-scheme architecture, Cu2O was protected against photocorrosion and the photocatalyst was stable for 42 h. In addition, the synergistic interactions between reduced titania and Cu2O resulted in 0.13% photoreduction of CO2 to CH4 [67]. Yin et al. grafted copper oxide nanoclusters onto niobate nanosheets, which served as a light harvesting component. This work found that photogenerated holes in the valence band of niobate nanosheets react with water and photogenerated electrons in the conduction band is injected into the copper oxide nanoclusters resulting in the production of carbon monoxide [68]. Lan et al. loaded Cu nanoparticles onto TiO2 which exhibited significantly higher photocatalytic activity than pure TiO2 for CO2 reduction and selective for carbon monoxide [69]. Bae et al. synthesized zinc oxide-copper (I) oxide hybrid nanoparticles in colloidal forms with copper (I) oxide nanocubes bound to zinc oxide. This photocatalyst exhibited high selectivity for methane (>99%) which was attributed to the optimal band alignment of zinc oxide and copper (I) oxide, surface defects, high surface area and colloidal morphology of the catalyst [70]. Nogueira et al. evaluated the effect of electrolyte on photochemical CO2 reduction using CuO nanoparticles and found electrolyte strongly influences product selectivity. NaOH led to methane production while Na2C2O4 led to carbon monoxide production and KBrO3 led to oxygen production [71]. Dedong et al. synthesized a Cu2O/coal-based carbon nanoparticle hybrid to reduce CO2 to methanol [72]. Zhang et al. fabricated an Ag-Cu2O/ZnO nanorod hybrid catalyst and found Cu2O enhances the CO2 chemisorption on the catalyst surface while the formation of Z-scheme system between Cu2O and ZnO facilitates charge separation. Silver nanoparticles onto Cu2O leads to higher photocatalytic activity due to electron transfer from silver to Cu2O [73]. Aguirre et al. found that TiO2 can protect Cu2O from undergoing photocorrosion when a Cu2O/TiO2 hybrid catalyst is formed. This hybrid creates an efficient Z-scheme which facilitates electron transfer from TiO2 to Cu2O [74]. Park et al. synthesized both CuO-TiO2 and Cu2O-TiO2 and demonstrated this hybrid catalyst has enhanced light absorption and rapid charge separation due to the intrinsic p-n heterojunction of the material, leading to improved photocatalytic activity [75]. Gusain et al. prepared reduced graphene oxide-copper oxide nanocomposites by covalent grafting of CuO nanorods onto reduced graphene oxide. Bare CuO showed low photocatalytic activity due to rapid charge carrier recombination but rGO-Cu2O and rGO-CuO showed significantly higher photocatalytic activity in the reduction of CO2 to methanol under visible light irradiation [76]. Zhai et al. synthesized a core-shell structure containing two co-catalysts capable of reducing CO2 to methane and carbon monoxide. They propose that the Cu2O shell provides sites for CO2 reduction while the Pt core extracts photogenerated electrons from TiO2 [77]. Wu et al. observed that the (110) facet of a single Cu2O particle is capable of reducing CO2 to methanol while the (100) facet is inert. The oxidation state of the active sites changes from Cu(I) to Cu(II) due to CO2 and H2O adsorption and changes back to Cu(I) after CO2 conversion under visible light irradiation [78]. Table 2 summarizes the different types of semiconductors used in photocatalytic CO2/bicarbonate reduction to formate and other products.

4.2. Phthalocyanine-Semiconductor Composites

Phthalocyanines are macrocyclic heterostructures with aromatic conjugated complexes consisting of nitrogen and carbon atoms. They offer enormous potential as the light harvesting components of dye-sensitized semiconductors [79]. They also offer high photo and thermal stability and low toxicity. In addition, the close position of the TiO2 conduction band with respect to the LUMO orbital energy of the phthalocyanines favors efficient charge transfer. When the phthalocyanines molecule is excited by visible light, electrons are excited from the HOMO to the LUMO energy level and is then injected into the conduction band of TiO2.
In 1997, Premkumar et al. was able to produce formate from carbon dioxide upon photo-reduction by metal phthalocyanines (cobalt and zinc) adsorbed onto a Nafion membrane that acted as a photocatalyst in acidic aqueous solution containing triethanolamine as a hole scavenger [80,81]. The catalyst was immobilized in the Nafion membrane and because they are physically separated from the solution phase they are able to exhibit high photocatalytic efficiency (Figure 8). It was found that the membranes behave as a p-type semiconductor and high turnover numbers were achieved. In 2009, Liu et al. synthesized CoPc-loaded TiO2 and in the presence of NaOH, produced high formate yield [82]. Na2SO3 was used as a hole scavenger and under optimal conditions, 1032 µmol/g cat of formate was produced. It was found that with the addition of CoPc, formate production significantly increased, with optimal CoPc loading being 0.5 wt%. The contact of TiO2 and CoPc involves redistribution of charge. Because the oxidation potential of S1 of CoPc is higher than the conduction band of TiO2, while the energy of T1 is lower than the conduction band of TiO2 (Figure 9), it is thermodynamically feasible for electron transfer from CoPc to TiO2. When this occurs, CoPc is oxidized to CoPc*+. The influence of NaOH, hole scavenger and irradiation time has been investigated. Because NaOH can dissolve more CO2 than pure water, making HCO3 the predominant form in aqueous solution therefore accelerating the photoreduction [82]. The mechanism of CO2 reduction with CoPc-loaded TiO2 begins with photo-excitation of the CoPc from S0 to S1, which can subsequently transfer its electrons to the conduction band of TiO2. CO2 molecules adsorbed onto the surface of TiO2 gain electrons and are reduced (Equations (10) and (11)) [81].
CoPc (S0) + hυ → CoPc*(S1)
CoPc*(S1) + TiO2(CB) → CoPc•+ + e(TiO2/CB).
In 2009, Zhao et al. also synthesized a CoPc/TiO2 nanocomposite to reduce CO2 under visible light irradiation [83]. CoPc molecules are excited first and are able to inject their electrons into TiO2, this allows for the increased separation of electron/hole pairs and thus increasing the photocatalytic efficiency to 1714 µmol/g cat. In 2013, Yazdanpour et al. synthesized copper phthalocyanine-modified TiO2 (CuPc/TiO2), which was coated on a stainless steel mesh and used for the photo-reduction of CO2 in the presence of visible light [84]. Mele et al. synthesized copper and zinc phthalocyanine-loaded TiO2 under visible light irradiation [79]. They reported that CuPc/TiO2 was the most efficient yielding formate production of 208.5 µmol/g cat. In 2015, Mele expanded on their work of CuPc-loaded TiO2 by studying both Cu(II) porphyrin and Cu(II) phthalocyanine loaded onto TiO2 [85]. Cu(II) phthalocyanine-loaded TiO2 was more efficient in the photoreduction of CO2 to formate due to its favorable reduction potential. This work reported a detailed study of the effect of the amount of catalyst, initial pH, copper loading in the CuPc-TiO2 composite, irradiation source and sensitizers. This work reported production values for formate of 239.5 µmol/g cat. Table 3 summarizes various phthalocyanine-enhanced semiconductors to reduce CO2/bicarbonate to formate and other products.

4.3. Metal-Organic Frameworks

Metal-organic frameworks (MOFs) are a class of crystalline materials made by joining metal ions or clusters and organic linkers together. Inorganic metals are called secondary building units (SBU) and organic units are ditopic or polytopic organic carboxylates. The different combinations of organic and inorganic components, as well as varying the geometry, size and functionality has led to more than 20,000 different MOFs being synthesized [86]. The three most distinguishing features of MOFs are their large surface area and ultrahigh porosity; as well as its permanent porosity. MOFs are stable structures that do not collapse upon the removal of solvent. Compared to commercial TiO2 (P-25) that have surface area of only 35–65 m2/g, MOFs have a tremendous amount of surface area. Surface area of MOFs typically range from 1000 to 10,000 m2/g and their porosity is greater than 50% of the MOF crystal volume [86]. MOFs are also shown to be thermally and chemically stable. It is these properties of MOFs that give them a great advantage over other porous materials such as zeolites and carbon-based materials.
Metal-organic frameworks (MOFs) are emerging as a new type of promising photocatalysts. MOFs are a class of crystalline and microporous material with a vast array of topologies and applications in numerous fields such as gas sensing, catalysis, gas storage and separation and drug delivery [87]. In MOFs the metal clusters can be regarded as inorganic semiconductor quantum dots that are linked by organic linkers serving as antennas to activate the quantum dots [88]. Electron injections occurs from the photo-excited organic linkers to the metal clusters, termed linker-to-cluster charge transfer (LCCT). MOFs are superior to semiconductors because their light absorption ability can be easily tuned by modifications on the organic linkers to make them visible-light responsive [89]. Table 4 summarizes CO2 reduction using various metal-organic frameworks.
Fu et al. synthesized a Ti-based MOF Ti8O8(OH)4(BDC)6 (where BCD = benzene-1,4-dicarboxylate) and the photocatalytic reduction of CO2 was performed in acetonitrile with triethanolamine (TEOA) as a sacrificial agent under visible light irradiation [90]. This MOF exhibited extremely high surface area of 1302 m2/g that produced 8.14 µmol of formate after 10 h. Upon irradiation in the LMCT band, a long-lived excited charge separate state occurs by transferring an electron from the organic ligand to Ti4+ (Figure 10). Sun et al. synthesized NH2-UiO-66(Zr), which is a Zr-containing MOF composed of hexameric Zr6O32 units linked by benzenedicarboxylate (BDC) [88]. This study substituted the BDC ligand with 2-aminoterephthalic acid (ATA), which rendered the MOF visible-light responsive due to the amino substituent. In the presence of triethanolamine as a sacrificial agent, visible light irradiation was able to reduce CO2 to formate reaching production values of 20.7 µmol/g cat-hr. Wang et al. reported synthesizing a series of earth-abundant Fe-containing MOFs that are able to reduce CO2 under visible light irradiation [89]. The direct excitation of the Fe-O clusters induces the electron transfer from O2− to Fe3+ to form Fe2+, which is responsible for the CO2 reduction. In addition, when these MOFs are functionalized with an amine substituent, their catalytic efficiency is greatly increased. This is attributed to the existence of dual excitation pathways: excitation of the NH2 functionality is followed by electron transfer to the Fe center and the direct excitation of Fe-O clusters. In the presence of TEOA as a sacrificial agent, formate production reached 178 µmol. Fei et al. incorporated a manganese bipyridine complex [Mn(bpydc)-(CO)3Br] into a robust Zr(IV)-based MOF for CO2 reduction to formate [91]. [Ru(dmb)3]2+ was used as a photosensitizer and 1-benzyl-1,4-dihydronicotinamide (BNAH) was used as a sacrificial agent (Figure 11). The enhanced photocatalytic efficiency was ascribed to the structure of the framework providing isolated active sites, which stabilize the catalyst. The catalyst maintained its crystallinity and was reused over several runs. Most recently, Lee et al. synthesized a mixed metal (Zr/Ti) MOF to reduce CO2 to formate under visible light irradiation [92]. TEOA was used as a sacrificial base and 1-benzyl-1,4-dihydronicotinamide (BNAH) as a sacrificial reductant. The MOFs were studied by photoluminescence spectroscopy to study charge transfer from organic linker to the metal cluster. This work resulted in high turnover numbers and the catalysts were recyclable over three runs.

5. Future Perspectives and Outlook

The majority of this review is focused on semiconductors for CO2/bicarbonate reduction because they are the most widely used photocatalyst. However, quantum efficiencies are still low due to extremely rapid electron/hole recombination within the semiconductor, narrow optical absorption of the semiconductor, scattering of the semiconductor, competing hydrogen evolution reaction and products getting re-oxidized to CO2. Though these issues can be mitigated by using photosensitizers and hole scavengers, a lot of work still needs to be done to make the reaction more efficient. In addition, semiconductors eventually degrade over time, so more durable catalysts need to be designed. Plasmon-enhanced photocatalysis is a rapidly expanding field due to the excellent optical properties of plasmonic noble-metal nanoparticles such as Au and Ag. Plasmonic nanoparticles primarily absorb in the visible region of the solar spectrum and are known to generate hot electrons that can be injected into the conduction band of semiconductors, allowing for excess electrons available for reduction. Plasmonic nanoparticles, along with organic dyes, can act as excellent photosensitizers for semiconductors.
Moving forward, design strategies for photocatalysts include the fabrication of three-dimensional hierarchical nanostructures to maximize surface area, catalyst kinks and defects and grain boundaries to allow for molecular adsorption to the catalyst surface. Different metal crystal structures and surface facets lead to different product distribution, therefore tuning the metal surface structure will allow for precise control over product selectivity. The addition of photosensitizers to render a catalyst visible light-responsive will maximize the region of optical absorption and the ability to utilize more of the solar spectrum. With these techniques, rational design of efficient, selective and durable catalysts is possible. With the development of these catalysts, one can then scale up and design reactors that can convert CO2 or flue gas into value-added products.

6. Conclusions

In conclusion, many significant achievements have been made in the field of photocatalytic CO2 reduction. In relation to hydrogen production, which remains an impactful renewable energy source, the case is made for the chemical conversion of CO2 to formate in terms of near term economic feasibility and in juxtaposition to geological storage and sequestration. Based on this review, one can see that the majority of photocatalysts for CO2 reduction to formate are comprised of semiconductors. However, due to their narrow region of optical absorption, many studies have incorporated ruthenium complexes, phthalocyanines, and employed metal-organic framework catalysts designed to be visible-light responsive. Other catalysts include the use of graphene and graphene oxide as electron sinks to improve charge separation. Albeit much effort has been put into studying CO2 reduction, the photoconversion efficiency and selectivity for desired products is still low. The mechanism of CO2 reduction still needs to be investigated in depth. The one-electron induced activation of CO2 into the anion radical CO2•− still requires a high overpotential, which seems to be the rate-limiting step. The development of photocatalytic materials for CO2 reduction is rapidly expanding and will continue to reach higher efficiencies. As apparent quantum efficiencies improve with current research trends and new investigators entering the field, we assert that more practical device and design will become a major research component to the success of greenhouse gas mitigation and regenerative energy.

Author Contributions

Conceptualization, H.P. and M.D.H.; Writing—Original Draft Preparation, H.P.; Writing—Review & Editing, H.P. and M.D.H.; Visualization, H.P.; Supervision, M.D.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors wish to thank NSF award # IIA-1301346, for support of the publications by HP and MDH included in this review article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barber, J. Photosynthetic energy conversion: Natural and artificial. Chem. Soc. Rev. 2009, 38, 185–196. [Google Scholar] [CrossRef] [PubMed]
  2. Sayigh, A. Comprehensive Renewable Energy; Elsevier: Amsterdam, The Netherlands, 2012. [Google Scholar]
  3. Hammarström, L.; Hammes-Schiffer, S. Artificial photosynthesis and solar fuels. Acc. Chem. Res. 2009, 42, 1859–1860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. British Petroleum (BP). Statistical Review of World Energy; BP plc: London, UK, 2020. [Google Scholar]
  5. Kannan, N.; Vakeesan, D. Solar energy for future world—A review. Renew. Sustain. Energy Rev. 2016, 62, 1092–1105. [Google Scholar] [CrossRef]
  6. Olah, G.A.; Goeppert, A.; Surya Prakash, G.K. Beyond Oil and Gas: The Methanol Economy; Wiley-VCH: Weinheim, Germany, 2006. [Google Scholar]
  7. Olah, G.A.; Mathew, T.; Goeppert, A.; Surya Prakash, G.K. Difference and significance of regenerative vs. renewable carbon fuels and products. Top. Catal. 2018, 61, 522–529. [Google Scholar] [CrossRef]
  8. Yue, D.; You, F.; Snyder, S.W. Biomass-to-bioenergy and biofuel supply chain optimization: Overview, key issues and challenges. Comput. Chem. Eng. 2014, 66, 36–56. [Google Scholar] [CrossRef]
  9. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic Reduction of CO2 on TiO2 and Other Semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef]
  10. Windle, C.D.; Perutz, R.N. Advances in molecular photocatalytic and electrocatalytic CO2 reduction. Coord. Chem. Rev. 2012, 256, 2562–2570. [Google Scholar] [CrossRef]
  11. Nitopi, S.; Bertheussen, E.; Scott, S.B.; Liu, X.; Engstfeld, A.K.; Horch, S.; Seger, B.; Stephens, I.E.L.; Chan, K.; Hahn, C.; et al. Progress and perspectives of electrochemical CO2 reduction on copper in aqueous electrolyte. Chem. Rev. 2019, 119, 7610–7672. [Google Scholar] [CrossRef] [Green Version]
  12. Lee, M.-Y.; Park, K.T.; Lee, W.; Lim, H.; Kwon, Y.; Kang, S. Current achievements and the future direction of electrochemical CO2 reduction: A short review. Crit. Rev. Environ. Sci. Technol. 2020, 50, 769–815. [Google Scholar] [CrossRef]
  13. Chang, X.; Wang, T.; Gong, J. CO2 photo-reduction: Insights into CO2 activation and reaction on surfaces of photocatalysts. Energy Environ. Sci. 2016, 9, 2177–2196. [Google Scholar] [CrossRef]
  14. Morris, A.J.; Meyer, G.J.; Fujita, E. Molecular approaches to the photocatalytic reduction of carbon dioxide for solar fuels. Acc. Chem. Res. 2009, 42, 1983–1994. [Google Scholar] [CrossRef] [PubMed]
  15. Hong, J.; Zhang, W.; Ren, J.; Xu, R. Photocatalytic reduction of CO2: A brief review on product analysis and systematic methods. Anal. Methods 2013, 5, 1086–1097. [Google Scholar] [CrossRef]
  16. Xie, S.; Zhang, Q.; Liu, G.; Wang, Y. Photocatalytic and photoelectrocatalytic reduction of CO2 using heterogeneous catalysts with controlled nanostructures. Chem. Commun. 2016, 52, 35–59. [Google Scholar] [CrossRef] [PubMed]
  17. Kumar, B.; Llorente, M.; Froehlich, J.; Dang, T.; Sathrum, A.; Kubiak, C.P. Photochemical and Photoelectrochemical Reduction of CO2. Annu. Rev. Phys. Chem. 2012, 63, 541–569. [Google Scholar] [CrossRef] [Green Version]
  18. Song, C. Global challenges and strategies for control, conversion and utilization of CO2 for sustainable development involving energy, catalysis, adsorption and chemical processing. Catal. Today 2006, 115, 2–32. [Google Scholar] [CrossRef]
  19. Janáky, C.; Hursán, D.; Endrődi, B.; Chanmanee, W.; Roy, D.; Liu, D.; de Tacconi, N.R.; Dennis, B.H.; Rajeshwar, K. Electro- and photoreduction of carbon dioxide: The twain shall meet at copper oxide/copper interfaces. ACS Energy Lett. 2016, 1, 332–338. [Google Scholar]
  20. Leonard, D.P.; Pan, H.; Heagy, M.D. Photocatalyzed reduction of bicarbonate to formate: Effect of ZnS crystal structure and positive hole scavenger. ACS Appl. Mater. Interfaces 2015, 7, 24543–24549. [Google Scholar] [CrossRef]
  21. Feaster, J.T.; Shi, G.; Cave, E.R.; Hatsukade, T.; Abram, D.N.; Kuhl, K.P.; Hahn, C.; Norskov, J.S.; Jaramillo, T.F. Understanding selectivity for the electrochemical reduction of carbon dioxide to formic acid and carbon monoxide on metal electrodes. ACS Catal. 2017, 7, 4822–4827. [Google Scholar] [CrossRef]
  22. Pang, R.; Teramura, K.; Asakura, H.; Hosokawa, S.; Tanaka, T. Role of bicarbonate ions in aqueous solution as a carbon source for photocatalytic conversion of CO2 into CO. ACS Appl. Energy Mater. 2019, 2, 5397–5405. [Google Scholar] [CrossRef]
  23. NOAA National Centers for Environmental Information. State of the Climate: Global Climate Report for June 2019; NCEI: Asheville, NC, USA, 2019. [Google Scholar]
  24. Harry-O’Kuru, R.E.; Biresaw, G.; Tisserat, B.; Evangelista, R. Synthesis of polyformate esters of vegetable oils: Milkweed, pennycress, and soy. J. Lipids 2016, 2016, 1–12. [Google Scholar] [CrossRef] [Green Version]
  25. Jouny, M.; Luc, W.; Jiao, F. General techno-economic analysis of CO2 electrolysis systems. Ind. Eng. Chem. Res. 2018, 57, 2165–2177. [Google Scholar] [CrossRef]
  26. Afshar, A.A.N. Chemical Profile: Formic Acid; TranTech Consultants, Inc.: Philadelphia, PA, USA, 2014. [Google Scholar]
  27. Bushuyev, O.; De Luna, P.; Dinh, C.T.; Tao, L.; Saur, G.; van de Lagemaat, J.; Kelley, S.O.; Sargent, E.H. What Should We Make with CO2 and How Can We Make It? Joule 2018, 2, 825–832. [Google Scholar] [CrossRef] [Green Version]
  28. Intergovernmental Panel on Climate Change. Climate Change 2014: Synthesis Report; IPCC: Geneva, Switzerland, 2014. [Google Scholar]
  29. British Petroleum (BP). Natural Gas, 67th ed.; BP plc: London, UK, 2018. [Google Scholar]
  30. Agarwal, A.S.; Zhai, Y.; Hill, D.; Sridhar, N. The electrochemical reduction of carbon dioxide to formate/formic acid: Engineering and economic feasibility. ChemSusChem 2011, 4, 1301–1310. [Google Scholar] [CrossRef] [PubMed]
  31. Spurgeon, J.M.; Kumar, B. A comparative technoeconomic analysis of pathways for commercial electrochemical CO2 reduction to liquid products. Energy Environ. Sci. 2018, 11, 1536–1551. [Google Scholar] [CrossRef]
  32. Rumayor, M.; Dominguez-Ramos, A.; Perez, P.; Irabien, A. A techno-economic evaluation approach to the electrochemical reduction of CO2 for formic acid manufacture. J. CO2 Util. 2019, 34, 490–499. [Google Scholar] [CrossRef]
  33. Halmann, M. Photoelectrochemical reduction of aqueous carbon dioxide on p-type gallium phosphide in liquid junction solar cells. Nature 1978, 275, 115. [Google Scholar] [CrossRef]
  34. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photoelectrocatalytic reduction of carbon dioxide in aqueous suspensions of semiconductor powders. Nature 1979, 277, 637. [Google Scholar] [CrossRef]
  35. Ulman, M.; Aurian-Blajeni, B.; Halmann, M. Photoassisted carbon dioxide reduction to organic compounds using rare earth doped barium titanate and lithium niobate as photoactive agents. Isr. J. Chem. 1982, 22, 177–179. [Google Scholar] [CrossRef]
  36. Aurian-Blajeni, B.; Halmann, M.; Manassen, J. Electrochemical measurement on the photoelectrochemical reduction of aqueous carbon dioxide on p-Gallium phosphide and p-Gallium arsenide semiconductor electrodes. Sol. Energy Mater. 1983, 8, 425–440. [Google Scholar] [CrossRef]
  37. Halmann, M.; Katzir, V.; Borgarello, E.; Kiwi, J. Photoassisted carbon dioxide reduction on aqueous suspensions of titanium dioxide. Sol. Energy Mater. 1984, 10, 85–91. [Google Scholar] [CrossRef]
  38. Inoue, H.; Moriwaki, H.; Maeda, K.; Yoneyama, H. Photoreduction of carbon dioxide using chalcogenide semiconductor microcrystals. J. Photochem. Photobiol. A Chem. 1995, 86, 191–196. [Google Scholar] [CrossRef]
  39. Kanemoto, M.; Shiragami, T.; Pac, C.; Yanagida, S. Semiconductor photocatalysis. 13. Effective photoreduction of carbon dioxide catalyzed by zinc sulfide quantum crystallites with low density of surface defects. J. Phys. Chem. 1992, 96, 3521–3526. [Google Scholar] [CrossRef]
  40. Johne, P.; Kisch, H. Photoreduction of carbon dioxide catalysed by free and supported zinc and cadmium sulphide powders. J. Photochem. Photobiol. A Chem. 1997, 111, 223–228. [Google Scholar] [CrossRef]
  41. Pan, H.; Chowdhury, S.; Premachandra, D.; Olguin, S.; Heagy, M.D. Semiconductor photocatalysis of bicarbonate to solar fuels: Formate production from copper(I) oxide. Acs. Sustain. Chem. Eng. 2018, 6, 1872–1880. [Google Scholar] [CrossRef]
  42. Pan, H.; Steiniger, A.; Heagy, M.D.; Chowdhury, S. Efficient production of formic acid by simultaneous photoreduction of bicarbonate and oxidation of glycerol on gold-TiO2 composite under solar light. J. CO2 Util. 2017, 22, 117–123. [Google Scholar] [CrossRef]
  43. Pan, H.; Heagy, M.D. Plasmon-enhanced photocatalysis: Ag/TiO2 nanocomposite for the photochemical reduction of bicarbonate to formic acid. Mrs. Adv. 2019, 4, 425–433. [Google Scholar] [CrossRef]
  44. Beierle, A.; Gieri, P.; Pan, H.; Heagy, M.D.; Manjavacas, A.; Chowdhury, S. Titanium nitride nanoparticles for the efficient photocatalysis of bicarbonate into formate. Sol. Energy Mater. Sol. Cells 2019, 200, 109967. [Google Scholar] [CrossRef]
  45. Baran, T.; Wojtyła, S.; Dibenedetto, A.; Aresta, M.; Macyk, W. Zinc sulfide functionalized with ruthenium nanoparticles for photocatalytic reduction of CO2. Appl. Catal. B Environ. 2015, 178, 170–176. [Google Scholar] [CrossRef]
  46. Kuwabata, S.; Nishida, K.; Tsuda, R.; Inoue, H.; Yoneyama, H. Photochemical reduction of carbon dioxide to methanol using ZnS microcrystallite as a photocatalyst in the presence of methanol dehydrogenase. J. Electrochem. Soc. 1994, 141, 1498–1503. [Google Scholar] [CrossRef]
  47. Irvine, J.T.S.; Eggins, B.R.; Grimshaw, J. Solar energy fixation of carbon dioxide via cadmium sulphide and other semiconductor photocatalysts. Sol. Energy 1990, 45, 27–33. [Google Scholar] [CrossRef]
  48. Eggins, B.R.; Robertson, P.K.J.; Murphy, E.P.; Woods, E.; Irvine, J.T.S. Factors affecting the photoelectrochemical fixation of carbon dioxide with semiconductor colloids. J. Photochem. Photobiol. A Chem. 1998, 118, 31–40. [Google Scholar] [CrossRef]
  49. Henglein, A.; Gutiérrez, M.; Fischer, C.H. Photochemistry of colloidal metal sulfides 6. Kinetics of interfacial reactions at ZnS-Particles. Ber. Der Bunsenges. Für Phys. Chem. 1984, 88, 170–175. [Google Scholar] [CrossRef]
  50. Kaneco, S.; Kurimoto, H.; Ohta, K.; Mizuno, T.; Saji, A. Photocatalytic reduction of CO2 using TiO2 powders in liquid CO2 medium. J. Photochem. Photobiol. A Chem. 1997, 109, 59–63. [Google Scholar] [CrossRef]
  51. Kaneco, S.; Shimizu, Y.; Ohta, K.; Mizuno, T. Photocatalytic reduction of high pressure carbon dioxide using TiO2 powders with a positive hole scavenger. J. Photochem. Photobiol. A Chem. 1998, 115, 223–226. [Google Scholar] [CrossRef]
  52. Kaneco, S.; Kurimoto, H.; Shimizu, Y.; Ohta, K.; Mizuno, T. Photocatalytic reduction of CO2 using TiO2 powders in supercritical fluid CO2. Energy 1999, 24, 21–30. [Google Scholar] [CrossRef]
  53. Liu, B.-J.; Torimoto, T.; Yoneyama, H. Photocatalytic reduction of CO2 using surface-modified CdS photocatalysts in organic solvents. J. Photochem. Photobiol. A Chem. 1998, 113, 93–97. [Google Scholar] [CrossRef]
  54. Kisch, H.; Lutz, P. Photoreduction of bicarbonate catalyzed by supported cadmium sulfide. Photochem. Photobiol. Sci. 2002, 1, 240–245. [Google Scholar] [CrossRef]
  55. Zhang, X.V.; Martin, S.T.; Friend, C.M.; Schoonen, M.A.A.; Holland, H.D. Mineral-assisted pathways in prebiotic synthesis:  Photoelectrochemical reduction of carbon(+IV) by manganese sulfide. J. Am. Chem. Soc. 2004, 126, 11247–11253. [Google Scholar] [CrossRef]
  56. Xia, X.-H.; Jia, Z.-J.; Yu, Y.; Liang, Y.; Wang, Z.; Ma, L.-L. Preparation of multi-walled carbon nanotube supported TiO2 and its photocatalytic activity in the reduction of CO2 with H2O. Carbon 2007, 45, 717–721. [Google Scholar] [CrossRef]
  57. Qin, G.; Zhang, Y.; Ke, X.; Tong, X.; Sun, Z.; Liang, M.; Xue, S. Photocatalytic reduction of carbon dioxide to formic acid, formaldehyde, and methanol using dye-sensitized TiO2 film. Appl. Catal. B Environ. 2013, 129, 599–605. [Google Scholar] [CrossRef] [Green Version]
  58. Ulagappan, N.; Frei, H. Mechanistic study of CO2 photoreduction in Ti silicalite molecular sieve by FT-IR spectroscopy. J. Phys. Chem. A 2000, 104, 7834–7839. [Google Scholar] [CrossRef]
  59. Sato, S.; Morikawa, T.; Saeki, S.; Kajino, T.; Motohiro, T. Visible-light-induced selective CO2 reduction utilizing a ruthenium complex electrocatalyst linked to a p-type nitrogen-doped Ta2O5 semiconductor. Angew. Chem. Int. Ed. 2010, 49, 5101–5105. [Google Scholar] [CrossRef] [PubMed]
  60. Sato, S.; Arai, T.; Morikawa, T.; Uemura, K.; Suzuki, T.M.; Tanaka, H.; Kajino, T. Selective co2 conversion to formate conjugated with H2O oxidation utilizing semiconductor/complex hybrid photocatalysts. J. Am. Chem. Soc. 2011, 133, 15240–15243. [Google Scholar] [CrossRef] [PubMed]
  61. Arai, T.; Tajima, S.; Sato, S.; Uemura, K.; Morikawa, T.; Kajino, T. Selective CO2 conversion to formate in water using a CZTS photocathode modified with a ruthenium complex polymer. Chem. Commun. 2011, 47, 12664–12666. [Google Scholar] [CrossRef]
  62. Suzuki, T.M.; Tanaka, H.; Morikawa, T.; Iwaki, M.; Sato, S.; Saeki, S.; Inoue, M.; Kajino, T.; Motohiro, T. Direct assembly synthesis of metal complex-semiconductor hybrid photocatalysts anchored by phosphonate for highly efficient CO2 reduction. Chem. Commun. 2011, 47, 8673–8675. [Google Scholar] [CrossRef] [Green Version]
  63. Iizuka, K.; Wato, T.; Miseki, Y.; Saito, K.; Kudo, A. Photocatalytic reduction of carbon dioxide over Ag cocatalyst-loaded ALa4Ti4O15 (A = Ca, Sr, and Ba) using water as a reducing reagent. J. Am. Chem. Soc. 2011, 133, 20863–20868. [Google Scholar] [CrossRef]
  64. Raja, K.S.; Smith, Y.R.; Kondamudi, N.; Manivannan, A.; Misra, M.; Subramanian, V. CO2 photoreduction in the liquid phase over Pd-supported on TiO2 nanotube and bismuth titanate photocatalysts. Electrochem. Solid State Lett. 2011, 14, F5–F8. [Google Scholar] [CrossRef]
  65. Mendoza, J.A.; Kim, H.K.; Park, H.K.; Park, K.Y. Photocatalytic reduction of carbon dioxide using Co3O4 nanoparticles under visible light irradiation. Korean J. Chem. Eng. 2012, 29, 1483–1486. [Google Scholar] [CrossRef]
  66. Qin, Z.; Tian, H.; Su, T.; Ji, H.; Guo, Z. Soft template inducted hydrothermal BiYO3 catalysts for enhanced formic acid formation from the photocatalytic reduction of carbon dioxide. Rsc Adv. 2016, 6, 52665–52673. [Google Scholar] [CrossRef]
  67. Ali, S.; Lee, J.; Kim, H.; Hwang, Y.; Razzaq, A.; Jung, J.-W.; Cho, C.-H.; In, S.-I. Sustained, photocatalytic CO2 reduction to CH4 in a continuous flow reactor by earth-abundant materials: Reduced titania-Cu2O Z-scheme heterostructures. Appl. Catal. B Environ. 2020, 279, 119344. [Google Scholar] [CrossRef]
  68. Yin, G.; Nishikawa, M.; Nosaka, Y.; Srinivasan, N.; Atarashi, D.; Sakai, E.; Miyauchi, M. photocatalytic carbon dioxide reduction by copper oxide nanocluster-grafted niobate nanosheets. ACS Nano 2015, 9, 2111–2119. [Google Scholar] [CrossRef] [PubMed]
  69. Lan, Y.; Xie, Y.; Chen, J.; Hu, Z.; Cui, D. Selective photocatalytic CO2 reduction on copper–titanium dioxide: A study of the relationship between CO production and H2 suppression. Chem. Commun. 2019, 55, 8068–8071. [Google Scholar] [CrossRef] [PubMed]
  70. Bae, K.-L.; Kim, J.; Lim, C.K.; Nam, K.M.; Song, H. Colloidal zinc oxide-copper(I) oxide nanocatalysts for selective aqueous photocatalytic carbon dioxide conversion into methane. Nat. Commun. 2017, 8, 1156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Nogueira, A.E.; Oliveira, J.A.; da Silva, G.T.S.T.; Ribeiro, C. Insights into the role of CuO in the CO2 photoreduction process. Sci. Rep. 2019, 9, 1316. [Google Scholar] [CrossRef] [Green Version]
  72. Dedong, Z.; Maimaiti, H.; Awati, A.; Yisilamu, G.; Fengchang, S.; Ming, W. Synthesis and photocatalytic CO2 reduction performance of Cu2O/Coal-based carbon nanoparticle composites. Chem. Phys. Lett. 2018, 700, 27–35. [Google Scholar] [CrossRef] [Green Version]
  73. Zhang, F.; Li, Y.-H.; Qi, M.-Y.; Tang, Z.-R.; Xu, Y.-J. Boosting the activity and stability of Ag-Cu2O/ZnO nanorods for photocatalytic CO2 reduction. Appl. Catal. B Environ. 2020, 268, 118380. [Google Scholar] [CrossRef]
  74. Aguirre, M.E.; Zhou, R.; Eugene, A.J.; Guzman, M.I.; Grela, M.A. Cu2O/TiO2 heterostructures for CO2 reduction through a direct Z-scheme: Protecting Cu2O from photocorrosion. Appl. Catal. B Environ. 2017, 217, 485–493. [Google Scholar] [CrossRef]
  75. Park, S.-M.; Razzaq, A.; Park, Y.H.; Sorcar, S.; Park, Y.; Grimes, C.A.; In, S.-I. Hybrid CuxO–TiO2 Heterostructured Composites for Photocatalytic CO2 Reduction into Methane Using Solar Irradiation: Sunlight into Fuel. ACS Omega 2016, 1, 868–875. [Google Scholar] [CrossRef] [Green Version]
  76. Gusain, R.; Kumar, P.; Sharma, O.P.; Jain, S.L.; Khatri, O.P. Reduced graphene oxide–CuO nanocomposites for photocatalytic conversion of CO2 into methanol under visible light irradiation. Appl. Catal. B Environ. 2016, 181, 352–362. [Google Scholar] [CrossRef]
  77. Zhai, Q.; Xie, S.; Fan, W.; Zhang, Q.; Wang, Y.; Deng, W.; Wang, Y. Photocatalytic Conversion of carbon dioxide with water into methane: Platinum and copper(I) oxide Co-catalysts with a core–shell structure. Angew. Chem. Int. Ed. 2013, 52, 5776–5779. [Google Scholar] [CrossRef]
  78. Wu, Y.A.; McNulty, I.; Liu, C.; Lau, K.C.; Liu, Q.; Paulikas, A.P.; Sun, C.-J.; Cai, Z.; Guest, J.R.; Ren, Y.; et al. Facet-dependent active sites of a single Cu2O particle photocatalyst for CO2 reduction to methanol. Nat. Energy 2019, 4, 957–968. [Google Scholar] [CrossRef]
  79. Mele, G.; Annese, C.; De Riccardis, A.; Fusco, C.; Palmisano, L.; Vasapollo, G.; D’Accolti, L. Turning lipophilic phthalocyanines/TiO2 composites into efficient photocatalysts for the conversion of CO2 into formic acid under UV–vis light irradiation. Appl. Catal. A Gen. 2014, 481, 169–172. [Google Scholar] [CrossRef] [Green Version]
  80. Premkumar, J.; Ramaraj, R. Photocatalytic reduction of carbon dioxide to formic acid at porphyrin and phthalocyanine adsorbed Nafion membranes. J. Photochem. Photobiol. A Chem. 1997, 110, 53–58. [Google Scholar] [CrossRef]
  81. Rajan Premkumar, J.; Ramaraj, R. Photoreduction of carbon dioxide by metal phthalocyanine adsorbed Nafion membrane. Chem. Commun. 1997, 4, 343–344. [Google Scholar] [CrossRef]
  82. Liu, S.; Zhao, Z.; Wang, Z. Photocatalytic reduction of carbon dioxide using sol-gel derived titania-supported CoPc catalysts. Photochem. Photobiol. Sci. 2007, 6, 695–700. [Google Scholar] [CrossRef]
  83. Zhao, Z.; Fan, J.; Xie, M.; Wang, Z. Photo-catalytic reduction of carbon dioxide with in-situ synthesized CoPc/TiO2 under visible light irradiation. J. Clean. Prod. 2009, 17, 1025–1029. [Google Scholar] [CrossRef]
  84. Yazdanpour, N.; Sharifnia, S. Photocatalytic conversion of greenhouse gases (CO2 and CH4) using copper phthalocyanine modified TiO2. Sol. Energy Mater. Sol. Cells 2013, 118, 1–8. [Google Scholar] [CrossRef]
  85. Mele, G.; Annese, C.; Accolti, L.; De Riccardis, A.; Fusco, C.; Palmisano, L.; Scarlino, A.; Vasapollo, G. Photoreduction of carbon dioxide to formic acid in aqueous suspension: A comparison between phthalocyanine/TiO2 and porphyrin/TiO2 catalysed processes. Molecules 2015, 20, 396–415. [Google Scholar] [CrossRef] [Green Version]
  86. Conner, W.C.; Falconer, J.L. Spillover in Heterogeneous Catalysis. Chem. Rev. 1995, 95, 759–788. [Google Scholar] [CrossRef]
  87. Gangu, K.K.; Maddila, S.; Mukkamala, S.B.; Jonnalagadda, S.B. A review on contemporary metal–organic framework materials. Inorg. Chim. Acta 2016, 446, 61–74. [Google Scholar] [CrossRef]
  88. Sun, D.; Fu, Y.; Liu, W.; Ye, L.; Wang, D.; Yang, L.; Fu, X.; Li, Z. Studies on photocatalytic CO2 reduction over NH2-Uio-66(Zr) and its derivatives: Towards a better understanding of photocatalysis on metal–organic frameworks. Chem. A Eur. J. 2013, 19, 14279–14285. [Google Scholar] [CrossRef] [PubMed]
  89. Wang, D.; Huang, R.; Liu, W.; Sun, D.; Li, Z. Fe-based MOFs for photocatalytic CO2 reduction: Role of coordination unsaturated sites and dual excitation pathways. Acs Catal. 2014, 4, 4254–4260. [Google Scholar] [CrossRef]
  90. Fu, Y.; Sun, D.; Chen, Y.; Huang, R.; Ding, Z.; Fu, X.; Li, Z. An amine-functionalized titanium metal–organic framework photocatalyst with visible-light-induced activity for CO2 reduction. Angew. Chem. Int. Ed. 2012, 51, 3364–3367. [Google Scholar] [CrossRef] [PubMed]
  91. Fei, H.; Sampson, M.D.; Lee, Y.; Kubiak, C.P.; Cohen, S.M. Photocatalytic CO2 reduction to formate using a Mn(I) molecular catalyst in a robust metal–organic framework. Inorg. Chem. 2015, 54, 6821–6828. [Google Scholar] [CrossRef] [PubMed]
  92. Lee, Y.; Kim, S.; Kang, J.K.; Cohen, S.M. Photocatalytic CO2 reduction by a mixed metal (Zr/Ti), mixed ligand metal-organic framework under visible light irradiation. Chem. Commun. 2015, 51, 5735–5738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Photo-excitation of a semiconductor and subsequent generation of charge carriers and decay pathways. X = electron donor, Y = electron acceptor.
Figure 1. Photo-excitation of a semiconductor and subsequent generation of charge carriers and decay pathways. X = electron donor, Y = electron acceptor.
Nanomaterials 10 02422 g001
Figure 2. Valence and conduction band edge potentials of various semiconductors, along with reduction potentials for CO2/HCOOH, NAD+/NADH, H+/H2 and H2O/O2.
Figure 2. Valence and conduction band edge potentials of various semiconductors, along with reduction potentials for CO2/HCOOH, NAD+/NADH, H+/H2 and H2O/O2.
Nanomaterials 10 02422 g002
Figure 3. Proposed mechanism for (a) Cu2O semiconductor and (b) Ag/Cu2O (blue block arrow represents resonant energy transfer from metal to semiconductor). Reproduced from [41], with permission from American Chemical Society, 2018.
Figure 3. Proposed mechanism for (a) Cu2O semiconductor and (b) Ag/Cu2O (blue block arrow represents resonant energy transfer from metal to semiconductor). Reproduced from [41], with permission from American Chemical Society, 2018.
Nanomaterials 10 02422 g003
Figure 4. Proposed mechanism of photocatalytic reduction of high-pressure CO2 using TiO2 and 2-propanol as a hole scavenger. Reproduced from [51], with permission from Elsevier, 1998.
Figure 4. Proposed mechanism of photocatalytic reduction of high-pressure CO2 using TiO2 and 2-propanol as a hole scavenger. Reproduced from [51], with permission from Elsevier, 1998.
Nanomaterials 10 02422 g004
Figure 5. Photocatalytic CO2 reduction to CH4, HCOOH and C2H5OH using TiO2-decorated multi-walled carbon nanotube (MWCNT).
Figure 5. Photocatalytic CO2 reduction to CH4, HCOOH and C2H5OH using TiO2-decorated multi-walled carbon nanotube (MWCNT).
Nanomaterials 10 02422 g005
Figure 6. Proposed mechanism for the photocatalytic CO2 reduction using bifunctionalized TiO2 film.
Figure 6. Proposed mechanism for the photocatalytic CO2 reduction using bifunctionalized TiO2 film.
Nanomaterials 10 02422 g006
Figure 7. Illustration of the Z-scheme for CO2 reduction. Reproduced from [60], with permission from American Chemical Society, 2011.
Figure 7. Illustration of the Z-scheme for CO2 reduction. Reproduced from [60], with permission from American Chemical Society, 2011.
Nanomaterials 10 02422 g007
Figure 8. Schematic illustration of the photocatalytic CO2 reduction at MPC adsorbed in a Nafion membrane. (Circle represents the semiconducting nature of the adsorbed MPC Nafion membrane. MPC = CoPC or ZnPC, TEA = triethanolamine, CB = conduction band, VB = valence band and EF = Fermi level).
Figure 8. Schematic illustration of the photocatalytic CO2 reduction at MPC adsorbed in a Nafion membrane. (Circle represents the semiconducting nature of the adsorbed MPC Nafion membrane. MPC = CoPC or ZnPC, TEA = triethanolamine, CB = conduction band, VB = valence band and EF = Fermi level).
Nanomaterials 10 02422 g008
Figure 9. Photocatalytic CO2 reduction using CoPc-loaded TiO2.
Figure 9. Photocatalytic CO2 reduction using CoPc-loaded TiO2.
Nanomaterials 10 02422 g009
Figure 10. Proposed mechanism for the photocatalytic CO2 reduction using NH2-MIL-125(Ti) under visible light irradiation.
Figure 10. Proposed mechanism for the photocatalytic CO2 reduction using NH2-MIL-125(Ti) under visible light irradiation.
Nanomaterials 10 02422 g010
Figure 11. Proposed mechanism for photocatalytic formate production using UiO-67-Mn(bpy)(CO)3Br.
Figure 11. Proposed mechanism for photocatalytic formate production using UiO-67-Mn(bpy)(CO)3Br.
Nanomaterials 10 02422 g011
Table 1. The main products of CO2 and water reduction and the corresponding reduction potentials with reference to NHE at pH 7 in aqueous solution [13,15].
Table 1. The main products of CO2 and water reduction and the corresponding reduction potentials with reference to NHE at pH 7 in aqueous solution [13,15].
Product ReactionE° (V vs. NHE)Equation
Hydrogen2H2O + 2e → 2OH + H2−0.41(1)
Methane CO2 + 8H+ + 8e → CH4 + 2H2O−0.24(2)
Carbon monoxide CO2 + 2H+ + 2e → CO + H2O−0.51(3)
Methanol CO2 + 6H+ + 6e → CH3OH + H2O−0.39(4)
Formic acid CO2 + 2H+ + 2e → HCOOH−0.58(5)
Ethane 2CO2 + 14H+ + 14e → C2H6 + 4H2O−0.27(6)
Ethanol 2CO2 + 12H+ + 12e → C2H5OH + 3H2O−0.33(7)
Oxalate 2CO2 + 2H+ + 2e → H2C2O4−0.87(8)
Table 2. Summary of CO2 reduction using semiconductors.
Table 2. Summary of CO2 reduction using semiconductors.
Light SourceReaction Medium and ElectrolyteCatalystFormate & Other ProductsRef.
Halogen lamp at 365 nmElectrode consisting of single crystal GaP in 0.05 M K2HPO4/KH2PO4GaPHCOOH: 1.2 × 10−2 M
HCHO: 3.2 × 10−4 M
CH3OH: 1.1 × 10−4 M
[33]
500 W Xe lampAqueous Suspension of semiconductor in water TiO2, ZnO, CdS, GaP, SiC, WO3TiO2: HCOOH: 1.8 x 10−3 M
CH3OH: 14.6 × 10−4 M
ZnO: HCOOH: 1.2 × 10−3 M
CH3OH: 3.5 × 10−4 M
CdS: HCOOH: 2.0 × 10−3 M
CH3OH: 11.7 × 10−4 M
GaP: HCOOH: 1.0 × 10−3 M
CH3OH: 11.0 × 10−4 M
SiC: HCOOH: 1.0 × 10−3 M
CH3OH: 53.5 × 10−4 M
WO3: HCOOH: 0
CH3OH: 0
[34]
High pressure Hg lampAqueous suspensions Doped BaTiO3, LiNbO3HCOOH and HCHO[35]
150 W Xe lampElectrodes in 0.5 M Na2CO3Single crystal p-GaP, p-GaAsp-GaP: HCOOH: 670 μmol
HCHO: 13 μmol
CH3OH: 10 μmol
p-GaAs: HCOOH: 320 μmol
HCHO: 5 μmol
[36]
High pressure Hg lampAqueous suspension in water RuO2-doped TiO2 HCOOH: 1.46 μmol/h
HCHO: 0.18 μmol/h
CH3OH: 0.2 μmol/h
[37]
High pressure Hg lampAqueous suspension in 1.5 × 10−3 M NaHCO3Cd-loaded ZnSQuantum efficiency:
HCOOH: 32.5%
HCHO: 42.0%
H2: 5.0%
[38]
High pressure Hg lampColloidal suspension in 0.7 M NaH2PO2ZnS HCOOH: 75.1 μmol/h
CO: 2.7 μmol/h
H2: 86.0 μmol/h
[39]
Tungsten-halogen lamp (λ > 350 nm)Aqueous suspension in water ZnS-loaded SiO2HCOOH: 10 mmol [40]
1000 W Xenon arc lampAqueous suspension in 0.2 M NaHCO3ZnSHCOOH: 140 mmol formate/g cat-hr[20]
1000 W Xenon arc lamp Aqueous suspension in 0.2 M NaHCO3Cu2OHCOOH: 2.78 mmol formate/g cat-hr[41]
1000 W Xenon arc lampAqueous suspension in 0.2 M NaHCO3Au/TiO2HCOOH: 55 mmol formate/g cat-hr[42]
1000 W Xenon arc lampAqueous suspension in 0.2 M NaHCO3Ag/TiO2HCOOH: 3.89 mmol formate/g cat-hr[43]
1000 W Xenon arc lampAqueous suspension in 0.2 M NaHCO3TiN/TiO2 HCOOH: 3000 mmol formate/g cat-hr[44]
150 W XBO arc lamp Aqueous suspension in water Ru nanoparticles-loaded ZnSHCOOH:0.006 M [45]
High pressure Hg lampColloidal suspension in H3PO4ZnS microcrystallite HCOOH: 11.6 μmol
HCHO: 11.8 μmol
CH3OH: 1.2 μmol
[46]
Medium pressure mercury lampAqueous suspension in water CdS, ZnO, SiC, Ba TiO3, SrTiO3HCOOH, HCHO[47]
Medium pressure mercury lampAqueous suspension in waterCdS, ZnSZnS: HCOOH: 320 μmol/L
CdS: HCOOH: 87 μmol/L
[48]
Medium pressure mercury lampAqueous suspension in water SiO2-stabilized ZnS HCOOH: 0.8% AQE[49]
990 W Xe lampAqueous suspension in waterTiO2HCOOH: 8 × 10−6 mol/g-cat[50]
Xe lampAqueous suspension in waterTiO2HCOOH: 2.3 × 10−6 mol/g-cat
CH4: 1.2 × 10−6 mol/g-cat
[51]
990 W Xe lampPower in supercritical fluid CO2TiO2HCOOH: 9 × 10−6 mol/g-cat[52]
500 W high pressure mercury arc lampAqueous suspension in water and 2-propanol CdSHCOOH: 0.4 μmol
HCHO: 1.9 μmol
CO: 0.8 μmol
H2: 0.5 μmol
[53]
100 W tungsten halogen lamp, 150 W xenon lampAqueous suspension in water CdS-SiO2, CdS-ZnSCdS-SiO2:
HCOOH: 10 × 10−5 M
HCHO: 3.4 × 10−5 M
CdS-ZnS:
HCOOH: 4.5 × 10−5 M
[54]
450 W medium pressure UV Hg arc lampAqueous suspension in water and 7.2 mM NaHS and 2.5 mM NaHCO3MnSHCOOH: 200 μM[55]
15 W UV lamp λ = 365 nmAqueous suspension in waterMWCNT-supported TiO2HCOOH: 125.1 μmol/g
CH4: 73.33 μmol/g
C2H5OH: 149.36 μmol/g
[56]
300 W Xe lampElectrode in water adjusted to pH 4Dye-sensitized TiO2 film HCOOH: 1.8 mmol/cm2
HCHO: 1.4 mmol/cm2
CH3OH: 1.9 mmol/cm2
[57]
Pulsed Nd:YAG laser at 10 Hz, λ = 266 nmMethanol Ti silicalite molecular sieve HCOOH, CO[58]
Xe lampAqueous suspension in MeCN/TEOAN-doped Ta2O5
(N-Ta2O5), linked with
electrocatalysts [Ru(dcbpy)(bpy)(CO)2]2+
or [Ru-
(dcbpy)2(CO)2]2+
HCOOH: 1.9% quantum yield [59]
Solar simulator with AM 1.5 filterAqueous suspension in water and 10 mM NaHCO3InP/Ru complex polymer hybrid HCOOH: 4.71 μmol/cm2[60]
Xe light source Aqueous suspension in waterCu2ZnSnS4HCOOH: 0.22 mM [61]
500 W Xe lampAqueous suspension in acetonitrile/triethanolamineN-doped Ta2O5 with [Ru(dcbpy)2(CO)2]2+ HCOOH: 1.9 quantum efficiency [62]
400 W high-pressure mercury lampAqueous suspension in waterAg-loaded ALa4Ti4O15 (A = Ca, Sr, Ba)BaLa4Ti4O15:
HCOOH: 150 μmol
[63]
300 W solar simulator with AM 1.5 filter Thin films submerged in 0.1 M H2SO4Anodized titanium oxide nanotubes (T-NT), Pd-decorated bismuth titanate (BTO)HCOOH: 160 μmol/h/g[64]
21 W LED lamp, λ = 510 to 620 nmAqueous suspension in waterCo3O4HCOOH: 4.53 µmol/g-h
HCHO: 0.62 µmol/g-h
[65]
300 W Xe lampAqueous suspension in 0.25 M NaOH and 0.1 M Na2SO3BiYO3HCOOH: 1.68 μmol/L[66]
100W solar simulator with AM 1.5 filterCO2 passing through solid photocatalystReduced titania-Cu2OCH4: 462 nmol/g[67]
Xe–Hg lampAqueous suspension in 0.5 M KHCO3Cu(II)-grafted Nb3O8 nanosheets CO: 1.5 µmol[68]
300 W Xe lampAqueous suspension in waterCu-TiO2CO: 244 μmol/g[69]
300 W Xe lampAqueous suspension in water and 0.2 M Na2CO3ZnO-Cu2O nanoparticlesCH4: 1080 μmol/gcat h
CO: 1.4 μmol/gcat h
[70]
5 W UVC lampAqueous suspension in 0.1 M Na2C2O4, 0.1 M KBrO3, 0.1 M NaOH or waterCuOCH4: 1000 µmol/L-g
CO: 6 µmol/L-g
[71]
300 W Xenon lampAqueous suspension in 1M NaOHCu2O/CNPsCH3OH: 236.43μmolg−1cat[72]
300 W Xe arc lampAqueous suspension in waterAg-Cu2O/ZnO nanorodsCO: 9.94 μmol/g[73]
1 kW high-pressure Hg (Xe) arc lampAqueous suspension in waterCu2O/TiO2CO: 2.11 μmol gcat−1 h−1[74]
100 W Xenon solar simulator with AM 1.5 filterPowder photocatalyst with CO2CuxO-TiO2CH4: 221.6 ppm/g-h[75]
20 W white cold LED flood lightAqueous suspension in DMF and waterrGO-CuO, rGO-Cu2OrGO-CuO:
CH3OH: 1228 μmol/g
rGO-Cu2O:
CH3OH: 862 μmol/g
[76]
200 W Xe lampPowder photocatalyst with CO2TiO2-loaded Pt and CuPt/TiO2:
CH4: 11 µmol/g-h
CO: 2.2 µmol/g-h
Cu/TiO2:
CH4: 8.7 µmol/g-h
CO: 5.4 µmol/g-h
Pt-Cu/TiO2:
CH4: 9.8 µmol/g-h
CO: 5.9 µmol/g-h
[77]
300 W Xe lampAqueous suspension in waterCu2OCH3OH: 1.2 mol g−1 h−1[78]
Table 3. Summary of CO2 reduction using phthalocyanines.
Table 3. Summary of CO2 reduction using phthalocyanines.
Light
Source
Reaction Medium and ElectrolyteCatalystFormate & Other ProductsRef.
300 UV-VIS UV-VIS lamp and 400 W Xe-Halogen lamp
S: SANOLIUX HRC uv vis lamp 300 W
H: RADIUM Xe-Halogen lamp 400 W
Aqueous suspension in water, pH adjusted by NaOH or H3PO4Lipophilic phthalocyanines/TiO2 compositesProduct yield in μmol/g cat
TiO2 with pH 3 and a S/H light source: 131 HCOOH
TiO2-H2Pc with pH 3 and a S/H light source: 75 HCOOH
TiO2-CuPc with pH 3 and a S/H light source: 208.5 HCOOH
TiO2-ZnPc with pH 3 and a S/H light source: 88.5 HCOOH
TiO2-CuPc with pH 7 and a S/H light source: 63.4 HCOOH
TiO2-CuPc with pH 13 and a S/H light source: 65.2 HCOOH
TiO2-CuPc with pH 3 and a S light source: 32.6 HCOOH
TiO2-CuPc with pH 3 and a H light source: 52.2 HCOOH
[79]
500 W tungsten-halogen lampMembrane dipped in water with 0.1 M TEA and 0.1 M HCIO4MP-Nafion, MPC-NafionNf/PP: 8 × 10−5 mol
Nf/CoTPP: 2.9 × 10−4 mol
Nf/FePC: 11 × 10−5 mol Nf/ZnPC: 2 × 10−4 mol
[80]
500 W tungsten–halogen lampMembrane dipped in 0.1 M triethanolamine and 0.1 M HClO4MPC-NafionNf-CoPc: 1.7 × 104 mol
Nf-ZnPc: 2.0 × 104 mol
[81]
500 W tungsten–halogen lampAqueous suspension in NaOHCoPc-TiO2 With 0 [HCHO]/M and 0 [CH3OH]/M:
HCOOH: 289.9 μmol/(g cat)
With 0 [HCHO]/M and 0.5 [CH3OH]/M:
HCOOH: 292.8 μmol/(g cat)
With 0 [HCHO]/M and 5 [CH3OH]/M:
HCOOH: 301.1 μmol/(g cat)
With 0.1 [HCHO]/M and 0 [CH3OH]/M:
HCOOH: 9731.3 μmol/(g cat)
With 1 [HCHO]/M and 0 [CH3OH]/M:
HCOOH: 82660.5 μmol/(g cat)
[82]
500 W tungsten-halogen lampAqueous suspension in 0.1 N NaOHCoPc-TiO21.0 wt% CoPc/TiO2:
HCOOH: 450.6 μmol/g cat
0.7 wt% In-situ CoPc/TiO2:
HCOOH: 1487.6 μmol/g cat
[83]
125 W high pressure mercury lampCO2 passed through catalyst-coasted reaction vessel CuPc-TiO214% photoconversion [84]
300 UV-VIS UV-VIS lamp and 400 W Xe-Halogen lampAqueous suspension in water, pH adjusted by NaOH or H3PO4CuPc-TiO2HCOOH: 239.5 μmol/gcat[85]
Table 4. Summary of CO2 reduction using metal-organic frameworks.
Table 4. Summary of CO2 reduction using metal-organic frameworks.
Light SourceReaction Medium and ElectrolyteCatalystFormate & Other ProductsRef.
500 W Xe lampAqueous suspension in 5:1 ratio of MeCN and TEOANH2-Uio-66(Zr)NH2-Uio-66(Zr): HCOOH: 13.2 μmol in 10 h
Mixed NH2-Uio-66(Zr): HCOOH:20.7 μmol in 10 h
NH2-UiO-66(Zr): HCOOH: None
Mixed NH2-UiO-66(Zr): HCOOH:7.28 μmol
[88]
300 W Xe lampAqueous suspension in 5:1 ratio of MeCN and TEOAMIL-101 (Fe),
MIL-53 (Fe),
MIL-88B (Fe),
NH2-MIL-101(Fe): HCOOH: 178 μmol
MIL-101(Fe): HCOOH: 59.0 μmol
NH2-MIL-53(Fe): HCOOH: 46.5 μmol
MIL-53(FE): HCOOH: 29.7 μmol
NH2-MIL-88(Fe): HCOOH: 30.0 μmol
MIL-88(Fe): HCOOH: 9.0 μmol
[89]
500 W Xe lampAqueous suspension in 5:1 ratio of MeCN and TEOANH2-MIL-125(Ti)NH2-MIL-125(Ti): HCOOH: 8.14 μmol
NH2-MIL-125(Ti) (λ > 450 nm): HCOOH: 3.83 μmol
[90]
470 nm LEDAqueous suspension in 4:1 ratio of DMF/TEOA solvent mixture containing 0.2 M 1-benzyl-1,4-dihydronicotiamide (BNAH)Mn(bpydc)(CO)3Br incorporated into Zr(IV)-based metal−organic frameworkTurnover numbers for Products:
UiO-67-Mn(bpy)(CO)3Br(b) for 4 h: HCOOH: 50
UiO-67-Mn(bpy)(CO)3Br(b) for 18 h: HCOOH: 110
[91]
300 W Xe arc lampAqueous suspension in 4:1 mixed solution of acetonitrile (MeCN)-triethanolamine (TEOA), which contained 1-benzyl-1,4-dihydronicotiamide (0.1 M, BNAH)Zr4.3Ti1.7O4(OH)4(C8H7O4N)5.17(C8H8O4N2)0.83Zr4.3Ti1.7O4(OH)4(C8H7O4N)5.17(C8H8O4N2)0.83:
HCOOH: 31.57 μmol
UiO-66(Zr/Ti)-NH2: HCOOH: 4.66 µmol
[92]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pan, H.; Heagy, M.D. Photons to Formate: A Review on Photocatalytic Reduction of CO2 to Formic Acid. Nanomaterials 2020, 10, 2422. https://doi.org/10.3390/nano10122422

AMA Style

Pan H, Heagy MD. Photons to Formate: A Review on Photocatalytic Reduction of CO2 to Formic Acid. Nanomaterials. 2020; 10(12):2422. https://doi.org/10.3390/nano10122422

Chicago/Turabian Style

Pan, Hanqing, and Michael D. Heagy. 2020. "Photons to Formate: A Review on Photocatalytic Reduction of CO2 to Formic Acid" Nanomaterials 10, no. 12: 2422. https://doi.org/10.3390/nano10122422

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