Next Article in Journal
Surface Display—An Alternative to Classic Enzyme Immobilization
Next Article in Special Issue
Environmental Applications of Photocatalytic Processes
Previous Article in Journal
Structure-Activity Relationship of Manganese Oxide Catalysts for the Catalytic Oxidation of (chloro)-VOCs
Previous Article in Special Issue
Fast Method for Testing the Photocatalytic Performance of Modified Gypsum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Gas Phase Photocatalytic CO2 Reduction, “A Brief Overview for Benchmarking”

1
Department of Energy Science & Engineering, DGIST, 333 Techno Jungang-daero, Hyeonpung-eup, Dalseong-gun, Daegu 42988, Korea
2
Department of Chemical Engineering, COMSATS University Islamabad, Lahore Campus, 1.5 KM Defence Road, Off Raiwind Road, Lahore 54000, Pakistan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2019, 9(9), 727; https://doi.org/10.3390/catal9090727
Submission received: 31 July 2019 / Revised: 19 August 2019 / Accepted: 19 August 2019 / Published: 28 August 2019
(This article belongs to the Special Issue Environmental Applications of Photocatalytic Processes)

Abstract

:
Photocatalytic CO2 reduction is emerging as an affordable route for abating its ever increasing concentration. For commercial scale applications, many constraints are still required to be addressed. A variety of research areas are explored, such as development of photocatalysts and photoreactors, reaction parameters and conditions, to resolve these bottlenecks. In general, the photocatalyst performance is mostly adjudged in terms of its ability to only produce hydrocarbon products, and other vital parameters such as light source, reaction parameters, and type of photoreactors used are not normally given appropriate attention. This makes a comprehensive comparison of photocatalytic performance quite unrealistic. Hence, probing the photocatalytic performance in terms of apparent quantum yield (AQY) with the consideration of certain process and experimental parameters is a more reasonable and prudent approach. The present brief review portrays the importance and impact of aforementioned parameters in the field of gas phase photocatalytic CO2 reduction.

Graphical Abstract

1. Introduction

Photocatalytic products, as a consequence of CO2 photoreduction, are industrially desirable with the additional benefit of normalizing anthropogenic CO2 [1]. It is inevitable to develop and design efficient photoreactors and optimize the photoreaction conditions in order to scale up the photocatalytic process, all in congruence with synthesis of robust photocatalysts [2]. Although there have been many studies pertaining to the synthesis of stable and efficient photocatalysts, only few studies are dedicated upon reaction engineering so as to ascertain the optimum reaction conditions and photo reactor design [3]. Both of these factors have significant influence on photocatalytic yield. For instance, product yield will be different for photoreactors with batch and continuous flow operation under different conditions of reaction parameters, feed type and concentration ratio, photocatalyst loadings, and sources of irradiation [4,5]. In particular, irradiation sources and its mode of irradiation over the photocatalyst have vital importance. As most of the frequently used photocatalysts are Ultraviolet (UV) light active, irradiation from UV source as compared to solar light will significantly enhance the product yield [6,7,8]. Further, solar concentrator technology with utilization of Fresnel lens for photocatalyst irradiation leads to enhanced light intensity and photon flux, which in turn renders a significant enhancement in yield of hydrocarbon products [8,9]. Although, most of the photocatalytic reactions are carried out at room temperature, concentrating solar light will increase the temperature of the system, which will induce the thermal effect that alters the yield [6,7].
Another decisive factor is the exposed area of the photocatalyst to interact with the light irradiations. Even for the different types of photocatalysts, when tested under similar conditions and with yields reported per gram of photocatalyst, a comparison can be misleading because the exposure area, per gram of the photocatalyst, to light will be different for powder catalysts and thin films [10,11]. Moreover, irrespective of photoreaction under standard conditions, yields have been reported in different customary units i.e., ppm cm−2 h−1, μmol g−1 h−1 and μmol cm−2 h−1 which further complicates the comparison [12,13,14,15]. Thus, photocatalysts that are tested under such different conditions and reporting consequent yield in different units lead to ambiguity for fair performance comparison. It is inappropriate to compare the activities of the photocatalysts on the basis of their intrinsic performance without considering effects of optimum conditions and reactor geometries because of these implications.
The aforementioned bottlenecks strongly advocate the necessity for standardized protocols in order to compare the results for photocatalytic CO2 reduction on equal grounds. Therefore, reporting yields on the basis of photonic efficiency, which incorporates the radiation source and exposure area of the photocatalyst, is an appealing approach [16]. This review specifically focuses on the parameters that influence the actual yield of the photocatalytic CO2 reduction reactions and outlines the standard testing practices for comparison and evaluation of performance. Moreover, we have calculated the apparent quantum yield (AQY) of different research works, on the basis of data available, for comparing the efficiencies of the photocatalysts.

2. Role of Organic Contaminations

Photocatalysts synthesis predominately involves organic materials as reaction reagents. The residues of such organic materials are not easily removed, even with calcination at higher temperatures. Photocatalytic CO2 reduction products mainly hydrocarbons, using such catalysts, originate concomitantly from these organic residues and CO2 as well [1,2,3]. Thus, yields from such photocatalysts are overestimated when compared to the actual yield originating from photocatalytic CO2 reduction. However, isotopic labelling by 13CO2 is carried out to rule out the possible involvement of these organics in overall photocatalytic yield. To understand the effects of organic contamination, Yang et al. performed Fourier Transform Infrared Spectroscopy (FTIR) investigations for photocatalytic CO2 reduction over Cu(I)/TiO2 photocatalyst synthesized by two different ways: one with polyethylene glycol (PEG) and another without PEG. The FTIR results showed that Cu(I)/TiO2 with PEG produces more CO (photocatalytic reduction product) as compared to Cu(I)/TiO2 without PEG as shown in Figure 1. This CO originates from organic contaminants even without the introduction of CO2. The possible mechanism of formation of CO is represented by Equations (1) and (2),
CO2 + C → 2CO
H2O + C → CO + H2
Calcination at high temperature and illumination under dry He/Ar is incapable of completely wiping out these contaminants [4,5]. However, UV treatment in the presence of H2O vapors is propitious in removing these contaminants [4]. Furthermore, they extended this study to elucidate the effects of prolonged and repeated pretreatment, for four cycles (7 h each) under He/H2O, for abolishing organic contaminants. In another study involving the synthesis of Ti-SBA-15 while using P123 (Pluronic acid) and TEOS (Tetraethylorthosilicate) by calcining at 550 °C for 6 h, they revealed the generation of significant amounts of hydrocarbon products (CH4, C2H6, C2H4) under He/H2O environment. However, the yield of such photocatalytic products was significantly decreased after the first cycle but not diminished completely as shown in Figure 2a–d [5]. Moreover, the control experiments for photocatalytic reduction of CO with H2O exhibited transition in enhanced selectivity from C1 to C2. Thus, CO produced as a consequence of organic contamination, as shown in Equations (1) and (2) has profound effects over yield and selectivity of photocatalytic CO2 reduction [5].
In a similar study by Busser et al., they reported the adsorption of CO2 over the surface of the photocatalyst during photodeposition of Cu/Cr upon Ga2O3 in the presence of CH3OH [1]. Their study confirmed that the CO2 originates from photooxidation of methanol as shown in Equation (3).
Cu2+ + CH3OH + H2O → Cu0 + CO2 + 6H+
Thus, the assessment of possible contribution of these organic contaminations is indispensable for gauging the actual yield.

3. Flow versus Batch Reactors

Photocatalysts that were tested under different reaction parameters and/or in different reactor geometries can have variability in reaction rate, yield and selectivity [6,7]. Literature suggests variety of reaction conditions and set-ups for photocatalytic CO2 reduction. Of these conditions, two type of reaction modes are extensively applied in photocatalytic CO2 reduction: (i) flow reaction system and (ii) batch reaction system [5,9,10,11,12]. However, the largely reported use of batch reactors obtains smaller yields of the photocatalytic products. Thus, the activity of the photocatalysts, tested under these different conditions, is imprecise to compare [2,10,13,14,15,16,17].
For batch reactors, it is difficult to understand and control the reaction mechanism and also product composition as well. This is because the products generated in the photocatalytic reaction may get re-adsorb over the surface of the photocatalysts or they can participate in side reactions to yield different products [9,18]. The key limitations of the batch reactor system include an accumulation of the products inside the reactor for a certain defined time which can lead to changes in the concentration of reactants by photocatalytic reactions itself, re-adsorption of the intermediate species or products, and the initiation of side reactions such as hydrogenation or re-oxidation to CO2. For instance, O2 that is produced during photocatalytic reduction when adsorbed on the surface of the photocatalyst might compete with CO2 for electron intake. Thus, batch reactors for CO2 photoreduction under various experimental conditions make it very difficult to compare the photocatalytic performance and they are not a suitable option for extended time and for industrial scale applications [12]. To resolve such issues, Pipelzadeh et al. designed a pressure swing reactor for facilitated product (CO mainly) desorption. In their configuration, the products were continuously recycled, thus periodic injection/evacuation of gases generated turbulence, which enhanced the CO production yield to 30–80%, depending upon flow rates as shown in Figure 3. Thus the continuous desorption of products is imperative which may surmount mass transfer limitations and the deactivation problems of the photocatalysts [19].
Contrary to batch reactors the re-adsorption of products and other aforementioned issues can be curtailed in continuous flow reactor system [9]. Despite that, the yield reported is still inadequate since these types of systems only allow for short residence time of reactants i.e., restricting reactants to make proper contact with photocatalysts. However, better performance can be obtained by optimizing the reaction conditions and using robust photocatalysts [20,21]. In another study, a twin reactor, as shown in Figure 4, was designed to avoid the possible re-oxidation of the photocatalytic products. The photocatalytic CO2 reduction results revealed that the production of hydrogen and oxygen in separate compartments and then use of as produced hydrogen in CO2 reduction increased the yield of the products [22].
As evident from Table 1, studies by In et al. clearly vindicate the efficacy of transition from batch to continuous flow reactor along with the use of robust photocatalysts. Their well-designed flow reaction system, as shown in Figure 5 consists of a mass flow controller (50 standard cm3, 20 °C, 1 atm) that regulates the flow of gases, a water bubbler (diameter = 3 cm, length = 15 cm) to maintain the desired humidity, a vacuum pump (pumping speed = 100 L min−1) to obtain high purity conditions by degassing the ultra-sealed photoreactor under a vacuum of 3.5 × 10−3 torr, and a gas chromatography unit for an automatic product intake and analysis (Shimadzu GC-2014, Restek Rt-Q Bond column, internal diameter = 0.53 mm, length = 30 m). Moreover, they used a ceramic porous disc (pore size = 1~1.6 µm) to support the catalyst and a quartz glass (diameter = 5 cm, thickness = 2 mm) to ensure efficient light transfer and the sealing of the reactor. Their research group also optimized the photoreactor dimensions to further improve the reaction process, (as shown in Figure 6a–c).

4. Reactor Geometry and Catalyst Support

Even the distribution of light over the surface of the photocatalyst is essential in reaping its full potential. In most of the reactor geometries where light impinges over the surface of the photocatalyst, from center or side, a shadow is casted on the opposite side. Consequently, a major portion of the photocatalyst is not activated. The contact of the light and photocatalyst might be enhanced by fabricating effective reactor geometries to achieve a uniform distribution of light and better photocatalyst dispersion [6,28,29]. Higher dispersion of the photocatalyst results in enhanced contact with reactants, better mass transfer and guarantees the maximum utilization of the light which all eventually translate to higher quantum yield [30]. To achieve this, a variety of approaches are reported in literature including the utilization of different reactor geometries and catalyst supports, as discussed in detail below.

4.1. Monolith Reactor

Tahir et al. studied the effect of photocatalyst dispersion by comparing performance of TiO2 coated micro channel monolith and cell type support (dispersed as single layer over stainless steel cell). Their study revealed a significant enhancement in CO production by TiO2 coated monolith as shown in Figure 7. This increment was mainly attributed to a broader exposed photocatalyst surface available for photocatalytic reaction [6]. In their study with gold-indium TiO2 dispersed over monolith, higher yields were reported particularly formation of C2 and C3 products. This enhancement was linked to effective utilization of photons owing to larger illuminated area of monolith [31]. A similar result of enhancement in CO2 reduction to CH4 was reported for montmorillonite modified TiO2 that was loaded over monolith as compared to the bare one [32].However, their performance is still restricted by limited light penetration despite the high flow rates, minimal pressure drop, and large surface area [33,34,35].

4.2. Fiber Optic Reactor

Fiber optic reactors are advantageous when compared to packed bed reactors owing to a better dispersion of photocatalyst and the spreading of light on large surface area [36]. Nguyen et al. compared the yield of the photocatalytic CO2 reduction carried over the photocatalyst coated on optical fiber and glass plate. Their study demonstrated ~15.2 time enhancement in CH4 and 11.6 times in C2H4 yield, for same amount of photocatalysts. This may be attributed to the synergistic effects of catalyst dispersion and effective light utilization [29]. Wang et al. also carried out CO2 photoreduction while using fiber optic reactor and they attributed the enhancement in yield to the gradual and uniform distribution of light upon irradiation [37]. Although optical fibers accompany the features of catalyst support and effective light distribution, they are saddled with the constraints of limited utilization of reactor volume and the shorter transportation distance of light from the tip of incidence. They occupy 20–30% of reactor volume, but a limited catalyst coated area restricts the effective use of incident light [33,35,37,38].

4.3. Monolith Fiber Optic Combined Reactor

Ola et al. combined the mutual effects of higher surface area of monolith and effective light distribution of fiber optics to fabricate internally illuminated monolith reactor, and compared the CO2 reduction performance of this system with slurry reactor. It was found that internal illumination, by optical fibers, of the monolith reactor enhanced quantum efficiency to 23 times owing to the higher surface area of monolith and even the distribution of light by optical fibers [39]. Liou et al. inserted carved polymethylmethacrylate (PMMA) made optical fibers into NiO/InTaO4 coated monolith (honey comb structure) as shown in Figure 8. This reactor when applied for photocatalytic CO2 reduction enhanced the yield of products (methanol and acetaldehyde). An enhancement in the yield can be attributed to increased surface area, higher photocatalyst loading and effective utilization of the light [35].

5. Light Irradiations

Light intensity and the type of irradiations are the most influential parameters for photocatalytic CO2 reduction [40]. Thus, to spur the solar chemical/fuel yielding reactions an effective contact, particularly at the microscopic level, between light and catalyst is imperative. As reported in the literature, a majority of the photocatalysts work efficiently in the UV range thus they are capable of only harnessing a limited range of solar spectrum [41]. To resolve this bottleneck, the modification of semiconductor is carried out for harvesting a wide range of solar spectrum. Other than that, light can be concentrated and channeled to obtain higher photon flux of irradiations [42]. Higher photon flux is not only conducive in propelling apparent quantum efficiency (AQE) but it also shores up the selectivity of multi-electron photoreactions to yield ethane and other long chain solar products. These claims were further vindicated by a study by Nagpal et al. while using TiO2 and CuInS (copper indium sulfide) nanocrystals as the photocatalyst. An enhanced AQY of 4.3% with a higher ethane selectivity of 70% was observed under concentrated sunlight owing to availability of a higher number of electrons [43]. Han et al. performed CO2 photoreduction by concentrating light over TiO2 and Pt/TiO2. The authors performed CO2 photoreduction experiments with different concentrating ratio (CR), which is defined as the ratio of concentrated light flux (amount of energy per unit time per unit area) on the photocatalyst surface to the ambient flux (under non-concentrated conditions). The light irradiation is concentrated and varied by changing the distance between the Fresnel lens (placed in between light source and photocatalyst) and photocatalyst surface, resulting in different light intensities with different light concentrated focal areas. Their study confirmed that the optimum concentration ratio (CR) significantly increased the AQY by 4.0 and 3.17 times for TiO2 and Pt/TiO2 respectively [42]. Similar results were reported by Li et al. where photocatalytic CH4 yield was improved by 29.5 and 6.2 times under optimum CR for untreated and pre-treated samples, respectively [44]. In their other study, they reported that CH4 yield, for photocatalytic CO2 reduction, over g-C3N4 at CR10 (10 times concentrating the light) was enhanced by factors of ~11.9 and ~16.0 for untreated and pre-treated samples respectively [45]. Tan et al. also reported enhanced AQY up to optimum light intensity but beyond that, it decreases as the number of photons exceeds the requirement for photocatalytic reaction. They also reported that reaction yields were significantly higher for AM 1.5 filter as compared to the UV cut-off filter which could be attributed to higher photon energy for UV leading to the generation of more photogenerated charges [46]. Based on these studies it can be gauged that better contact of light with photocatalyst under optimally concentrated light may lead to an exceptional yield increase.

6. Temperature

Concentrating solar light also increases the temperature depending upon the CR, due to long wavelength irradiations [42,47,48]. Photocatalytic CO2 reduction at an elevated temperatures is promising as it overcomes the thermal barriers that lead to slow reaction rate and marginal yields [48]. The efficacy of temperature rise for photoreaction can be underscored on the basis of enhanced effective collisions among photogenerated charges and reactants that directly relate to the reaction rate [30,46]. Furthermore, elevated temperature also facilitates the desorption of the products providing the way for the adsorption of the CO2 on vacant sites leading to increased reaction rate [49,50]. Wang et al. found that production rate almost doubled when the temperature was increased from 25 °C to 75 °C [37]. Desorption of methanol from the photocatalyst surface increases due to increase in temperature, thus providing more active sites for CO2 photoreduction. The AQY in turn also increased due to more efficient utilization of the incident light and enhanced CO2 adsorption as result of methanol desorption. However, the reaction temperature should not increase too much as it might also desorb the CO2 thus decelerating the photoreduction process. Similarly, Guan et al. attained a temperature up to 590 K for photocatalytic reduction of CO2 and they found temperature rise to be an effective factor for enhancing the yield of solar products [47]. They suggested that such an enhancement was due to the increased collision frequency of photons and diffusion rate of CO2 towards the surface active sites. Hence the thermal energy along with light irradiation can significantly improve the solar products yield by efficiently overcoming the kinetic barrier for CO2 photoreduction reactions. In another study Alxneit et al. revealed the rate of CH4 formation via photocatalytic CO2 reduction increase when the temperature was increased from 25 to 200 °C. Upon further increasing the temperature, the reaction rate decreased due to desorption of the reactants. However, with the aforementioned temperature range, the reaction rate increases with increasing temperature which indicates the importance of thermal step tending to decrease the surface coverage from intermediate species and products. These results clearly suggest that temperature dependence is one of the pivotal factors that influences the photocatalytic reaction rate and thereof the product yield [48]. In another study, Tahir et al. also confirmed the efficacy of temperature rise in photocatalytic CO2 reduction for enhanced CO production over In/TiO2 as shown in the Figure 9 [6]. Such an observation was again explained on the basis of adsorption-desorption phenomenon. Upon increasing the reaction temperature, the mass transfer of CO2 on the active sites is increased, which leads to increased CO2 adsorption, resulting in increased reaction rates. They also observed the raising temperature also transformed the product selectivity towards longer chain hydrocarbons. Similarly Zhang et al. reported, when temperature increased from 323 K to 343 K, the yield of the photocatalytic reaction became twice [50]. They also suggested that enhancement in the product yield is due to desorption of products on the photocatalyst surface, providing more chances of collisions between the excited states and adsorbed reactants. Although raising temperature is an effective strategy to obtain higher yields, different studies suggest that there is always an optimum temperature range to get the best AQY [42].

7. Effect of H2O/CO2 Feed Ratio

An established mixture of water vapors and CO2 gas (H2O/CO2) is considered to be a cost effective and invulnerable feed for photocatalytic reduction into chemicals/fuels. The feed ratio is another crucial factor with a profound impact over reaction rate and product yield. In addition, the affinity of the photocatalyst for H2O/CO2 may also lead to well tuning of the product selectivity. Yamashita et al. found highly selective methanol formation over the surface of hydrophilic Ti-Beta(OH) zeolites when compared to conventional Ti-Beta(F) zeolites [51]. Tahir et al. studied the effect of varying H2O/CO2 ratio by manipulating the CO2 flow rates. In their study they reported that at lower concentration of CO2, water could adsorb over the photocatalyst to react efficiently and give better yields. However, at higher CO2 concentrations, H2O has to compete for adsorption which may influence the yield. The same authors also studied the variation in H2O/CO2 ratio and its influence on the yield [6]. Zhang et al. also reported similar results mentioning enhancement in yield with increasing H2O/CO2 ratio [50]. In another study by Tahir et al., which reports enhancement in CH4 yield with increasing H2O/CO2 ratio. This is attributed to adsorption of excess water molecules over the photocatalyst and resulting in incremented ability to reduce CO2. Further increase in CO2 decelerated the yield, shown in Figure 10, due to competition between water and CO2 molecules on the photocatalyst active sites [30]. As the H2O/CO2 feed ratio is increased more water molecules will cover the photocatalyst surface due to hydrophilic nature of the material, which competes with the CO2 molecules to get adsorb on the photocatalyst active sites during the photoreduction process. Therefore, an optimum feed ratio is required for moderate adsorption of both water and CO2 molecules which in turn lead to the maximum CH4 yield. The importance of optimum H2O/CO2 ratio was further vindicated by Wu et al. in which they reported optimum H2O/CO2 ratio is essential for enhanced yield of methanol [52]. The investigation of Tan et al. also stressed the need for adsorption of optimum number of molecules of CO2 and H2O over the catalyst surface to obtain enhanced yields. Moreover it is important to note that, enhanced adsorption of any reactant will certainly hinder the adsorption of others [46]. Thus, a tradeoff should exist between CO2 and H2O to mitigate their competitive adsorption over the surface of the photocatalyst. Thus, due to rivalry in adsorption, an optimum feed ratio is indispensable to get improved yield.

8. Other Factors

In addition to the above mentioned parameters, there exists several other parameters which also alters the photocatalytic yield. A reported work investigated the performance of photocatalyst in cell type and monolith photoreactor which revealed that broader exposed area of photocatalyst, as for monolith catalyst support, improves the yield (shown in Figure 11). This can be related to enhancement in the illumination area of photocatalyst i.e., exposure of same weight of the catalyst in a larger area [29,30,39,42,53]. In addition to the conditions of high purity, ultra-high vacuum also ensures the absence of external impurities [9]. Moreover, by improving the surface properties like CO2 adsorption capability, photocatalytic yield can also be improved [44,45,54,55,56,57,58].

9. Benchmarking for Performance Evaluation

Based on the above discussion it can be well acknowledged that reactor geometries, reaction parameters, light intensity and wavelength are important parameters that play a pivotal role to control the product yields and selectivity. Thus photocatalysts that were tested under different conditions cannot be compared on equal grounds [42]. Like, Tan et al. reported that a maximum yield of 3.14 µmol g−1 was obtained at a light intensity of 177.2 mW cm−2 but in terms of AQY, the performance is better at 81 mW cm−2, as shown in Figure 12, for a given amount of photocatalyst [46]. This clearly reveals that although the photocatalyst is producing maximum yield at 177.2 mW cm−2 it still underperforms in capitalizing photogenerated charges. Thus, reporting performance of the photocatalytic reaction only on the basis of intrinsic capability of photocatalyst to yield the products is not a rationalized approach. On the contrary, evaluating the performance of the photocatalyst in terms of AQY, which incorporates reactor area, incident and harvested light, can be a more adequate approach [59]. For meaningful performance comparison, AQYs of the different photocatalysts are calculated and presented in Table 1, Table 2 and Table 3, using Equations (4)–(8) [11,59].
AQY   ( % )   =   number   of   reacted   electrons effective   number   of   incident   photons ×   100 % ,
number   of   reacted   electrons   = [ mole   of   product   produced   in   time ,   t ] × [ number   of   electrons   required   to   produce   1   mol   of   product ] ×   N A ,
effective   number   of   incident   photons   =   light   absorbed   by   the   photocatalyst average   photon   energy ×   t ,
light   absorbed   by   the   photocatalyst   =   H   ×   A ,
average   photon   energy   =   hc λ ,
where H is the apparent light input (W m−2), A is the geometric irradiation area (m2), h is Planck’s constant (6.626 × 10−34 J·s), c is speed of light ( 3 × 10 8 m s−1), λ is the average wavelength of light source (nm) and NA is Avogadro’s number ( 6.022 × 10 23 atoms mol−1).
Values for the calculations of AQYs included in this review are given in detail in Supplementary Tables S1 and S2. Moreover, prior to photocatalytic testing under CO2, repeated photocatalytic reduction tests with H2O and inert gas can rule out the possible role of organic contamination. Further confirmation of the carbon source of photocatalytic products can be ascertained by an isotopic labeling test while using 13CO2. Additionally, optimization of the reaction parameters, condition of high purity and effective photocatalyst and light contact can provide an even better judgment of the performance [1,4,9,60].

10. Conclusions

Photocatalytic CO2 reduction is a fascinating approach owing to its two-fold benefits of CO2 abatement and its subsequent conversion to renewable fuels/chemicals. However, optimizing this process has been a challenge for researchers as a variety of process parameters and reaction conditions indistinctly influence the product yield. One parameter is the reaction temperature which by increasing overcomes the kinetic barrier; but on the contrary, excessive temperature might also lead to decomposition. The reactants feed ratio also affects the product yield: at an optimum value, it synergizes the product yield and selectivity but beyond which, the non-stoichiometric ratios of H2O/CO2 are competing to get adsorbed on the photocatalyst, resulting in decreased productivity. In addition, optimal light intensity is imperative to produce stoichiometric amount of photogenerated charges for efficient transformation of adsorbed H2O/CO2 to solar fuel/chemicals; surplus of these charges, produced at higher light intensities, will recombine and eventually reduce the AQY. Moreover, productivity enhancement also depends on the choice of reactor type: continuous flow reactors have shown potential in overcoming the low yield, re-adsorption, and decomposition of photocatalytic products accustomed to batch reactors. Additionally, the reactor geometry with better light and photocatalyst contact notably affects the production yield. Knowing how these parameters affect the product yield, performance of the photocatalysts tested under different conditions are inappropriate to compare in frequently used customary units i.e., μmol g−1 or ppm cm−2. Hence, with the consideration of various process parameters and reaction conditions i.e., reaction temperature, feed ratio, irradiation source, reactor type and geometry, reporting yield of the photocatalysts in terms of AQYs rather than production rate is a more appropriate and pragmatic approach. Additionally, the removal of organic contamination, isotopic labelling, and photocatalytic CO2 reduction under inert and ultra-sealed condition can provide more realistic assessment of performance.
The overall aim of the present review article is to highlight the influence of key process and reaction parameters on CO2 photoreduction process to be used for benchmarking. While considering all the aforementioned parameters, a continuous and ultra-sealed gas phase reactor purged under high vacuum with an optimum blend of temperature, feed ratio, and offering larger contact between light and photocatalyst can be an effective approach for achieving maximum yield with the effective utilization of incident photons. Moreover, in order to compare the photocatalytic system performance, a unified parameter should be recognized; reporting yield in terms of AQY standardizes the photocatalytic performance analysis by normalizing the effect of the most influential parameters. Our aforementioned photoreaction setup and studies, well matching with the concluded benchmarking criteria, can be followed for future studies in the field of gas phase CO2 photoreduction.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/9/727/s1. Table S1: Computation of AQY performance of photocatalysts by In et al., Table S2: Computation of AQY performance of photocatalysts from other research works.

Author Contributions

S.-I.I. conceptualized and edited the manuscript, S.A., M.C.F., A.R., S.S., C.B.H., H.R.K., Y.H.P., Y.H., H.S.K., H.K., E.H.G., J.L. and D.K. wrote the manuscript.

Funding

The authors gratefully acknowledge the support of the Ministry of Science and ICT (2017R1E1A1A01074890 and 2017M2A2A6A01070912). This research was also supported by the Technology Development Program to Solve Climate Changes of the National Research Foundation (NRF) funded by the Ministry of Science and ICT (2015M1A2A2074670) as well as by the DGIST R&D Program of the Ministry of Science and ICT (19-BD-0404) and supported by a grant of the Korea Health Technology R&D Project through the Korea Health Industry Development Institute (KHIDI), funded by the Ministry of Health & Welfare, Republic of Korea (HI19C0506) and supported by Flux Photon Corporation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Busser, G.W.; Mei, B.; Pougin, A.; Strunk, J.; Gutkowski, R.; Schuhmann, W.; Willinger, M.; Schlögl, R.; Muhler, M. Photodeposition of Copper and Chromia on Gallium Oxide: The Role of Co-Catalysts in Photocatalytic Water Splitting. ChemSusChem 2014, 7, 1030–1034. [Google Scholar] [CrossRef] [PubMed]
  2. Razzaq, A.; Grimes, C.A.; In, S.-I. Facile Fabrication of a Noble Metal-free Photocatalyst: TiO2 Nanotube Arrays Covered with Reduced Graphene Oxide. Carbon 2016, 98, 537–544. [Google Scholar] [CrossRef]
  3. Shi, R.; Chen, Y. Controlled Formation of Defective Shell on TiO2 (001) Facets for Enhanced Photocatalytic CO2 Reduction. ChemCatChem 2019, 11, 2270–2276. [Google Scholar] [CrossRef]
  4. Yang, C.C.; Yu, Y.H.; Van Der Linden, B.; Wu, J.C.S.; Mul, G. Artificial Photosynthesis over Crystalline TiO2-Based Catalysts: Fact or Fiction? J. Am. Chem. Soc. 2010, 132, 8398–8406. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, C.C.; Vernimmen, J.; Meynen, V.; Cool, P.; Mul, G. Mechanistic Study of Hydrocarbon Formation in Photocatalytic CO2 Reduction over Ti-SBA-15. J. Catal. 2011, 284, 1–8. [Google Scholar] [CrossRef]
  6. Tahir, M.; Amin, N.S. Photocatalytic CO2 Reduction and Kinetic Study Over In/TiO2 Nanoparticles Supported Microchannel Monolith Photoreactor. Appl. Catal. A Gen. 2013, 467, 483–496. [Google Scholar] [CrossRef]
  7. Delavari, S.; Amin, N.A.S. Photocatalytic Conversion of CO2 And CH4 Over Immobilized Titania Nanoparticles Coated on Mesh: Optimization and Kinetic Study. Appl. Energy 2016, 162, 1171–1185. [Google Scholar] [CrossRef]
  8. Sorcar, S.; Thompson, J.; Hwang, Y.; Park, Y.H.; Majima, T.; Grimes, C.A.; Durrant, J.R.; In, S.-I. High-Rate Solar-Light Photoconversion of CO2 to Fuel: Controllable Transformation from C1 to C2 Products. Energy Environ. Sci. 2018, 11, 3183–3193. [Google Scholar] [CrossRef]
  9. Dilla, M.; Schlögl, R.; Strunk, J. Photocatalytic CO2 Reduction under Continuous Flow High-Purity Conditions: Quantitative Evaluation of CH4 Formation in the Steady-State. ChemCatChem 2017, 9, 696–704. [Google Scholar] [CrossRef]
  10. Kondratenko, E.V.; Mul, G.; Baltrusaitis, J.; Larrazábal, G.O.; Pérez-Ramírez, J. Status and Perspectives of CO2 Conversion into Fuels and Chemicals by Catalytic, Photocatalytic and Electrocatalytic Processes. Energy Environ. Sci. 2013, 6, 3112–3135. [Google Scholar] [CrossRef]
  11. Sorcar, S.; Hwang, Y.; Grimes, C.A.; In, S.-I. Highly Enhanced and Stable Activity of Defect-Induced Titania Nanoparticles for Solar Light-Driven CO2 Reduction into CH4. Mater. Today 2017, 20, 507–515. [Google Scholar] [CrossRef]
  12. Li, Y.; Wang, W.N.; Zhan, Z.; Woo, M.H.; Wu, C.Y.; Biswas, P. Photocatalytic Reduction of CO2 with H2O on Mesoporous Silica Supported Cu/Tio2 Catalysts. Appl. Catal. B Environ. 2010, 100, 386–392. [Google Scholar] [CrossRef]
  13. Zubair, M.; Kim, H.; Razzaq, A.; Grimes, C.A.; In, S.-I. Solar Spectrum Photocatalytic Conversion of CO2 to CH4 Utilizing TiO2 Nanotube Arrays Embedded with Graphene Quantum Dots. J. CO2 Util. 2018, 26, 70–79. [Google Scholar] [CrossRef]
  14. Zubair, M.; Razzaq, A.; Grimes, C.A.; In, S.-I. Cu2ZnSnS4 (CZTS)-ZnO: A Noble Metal-Free Hybrid Z-Scheme Photocatalyst For Enhanced Solar-Spectrum Photocatalytic Conversion of CO2 To CH4. J. CO2 Util. 2017, 20, 301–311. [Google Scholar] [CrossRef]
  15. Razzaq, A.; Sinhamahapatra, A.; Kang, T.H.; Grimes, C.A.; Yu, J.S.; In, S.-I. Efficient Solar Light Photoreduction of CO2 to Hydrocarbon Fuels Via Magnesiothermally Reduced TiO2 Photocatalyst. Appl. Catal. B Environ. 2017, 215, 28–35. [Google Scholar] [CrossRef]
  16. Kim, H.R.; Razzaq, A.; Grimes, C.A.; In, S.-I. Heterojunction pnp Cu2O/S-TiO2/CuO: Synthesis and Application to Photocatalytic Conversion of CO2 to Methane. J. CO2 Util. 2017, 20, 91–96. [Google Scholar] [CrossRef]
  17. Parayil, S.K.; Razzaq, A.; In, S.-I. Formation of Titania-Silica Mixed Oxides in Solvent Mixtures and Their Influences for the Photocatalytic CO2 Conversion to Hydrocarbon. J. Nanosci. Nanotechnol. 2015, 15, 7285–7292. [Google Scholar] [CrossRef] [PubMed]
  18. In, S.-I.; Nielsen, M.G.; Vesborg, P.C.K.; Hou, Y.; Abrams, B.L.; Henriksen, T.R.; Hansen, O.; Chorkendorff, I. Photocatalytic methane decomposition over vertically aligned transparent TiO2 nanotube arrays. Chem. Commun. 2011, 47, 2613–2615. [Google Scholar] [CrossRef] [PubMed]
  19. Pipelzadeh, E.; Rudolph, V.; Hanson, G.; Noble, C.; Wang, L. Applied Catalysis B: Environmental Photoreduction of CO2 on ZIF-8/TiO2 nanocomposites in a gaseous photoreactor under pressure swing. Appl. Catal. B Environ. 2017, 218, 672–678. [Google Scholar] [CrossRef]
  20. Adachi, K.; Ohta, K.; Mizuno, T. Photocatalytic reduction of carbon dioxide to hydrocarbon using copper-loaded titanium dioxide. Sol. Energy 1994, 53, 187–190. [Google Scholar] [CrossRef]
  21. Teraoka, Y.; Incorporated, F.; Suminoeku, N. Photocatalytic Reduction of CO2 with H2O on TiO2 And Cu/TiO2 Catalysts. Res. Chem. Intermed. 1994, 20, 815–823. [Google Scholar]
  22. Xiong, Z.; Kuang, C.C.; Lin, K.Y.; Lei, Z.; Chen, X.; Gong, B.; Yang, J.; Zhao, Y.; Zhang, J.; Xia, B.; et al. Enhanced CO2 photocatalytic reduction through simultaneously accelerated H2 evolution and CO2 hydrogenation in a twin photoreactor. J. CO2 Util. 2018, 24, 500–508. [Google Scholar] [CrossRef]
  23. Kim, H.R.; Razzaq, A.; Heo, H.J.; In, S.-I. Photocatalytic conversion of CO2 into hydrocarbon fuels with standard titania (Degussa P25) using newly installed experimental setup. Rapid Commun. Photosci. 2014, 2, 64–66. [Google Scholar] [CrossRef]
  24. Kim, K.; Razzaq, A.; Sorcar, S.; Park, Y.; Grimes, C.A.; In, S.-I. Hybrid mesoporous Cu2ZnSnS4 (CZTS)–TiO2 photocatalyst for efficient photocatalytic conversion of CO2 into CH4 under solar irradiation. RSC Adv. 2016, 6, 38964–38971. [Google Scholar] [CrossRef]
  25. 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]
  26. Parayil, S.K.; Razzaq, A.; Park, S.M.; Kim, H.R.; Grimes, C.A.; In, S.-I. Photocatalytic conversion of CO2 to hydrocarbon fuel using carbon and nitrogen co-doped sodium titanate nanotubes. Appl. Catal. A Gen. 2015, 498, 205–213. [Google Scholar] [CrossRef]
  27. Sorcar, S.; Hwang, Y.; Lee, J.; Kim, H.; Grimes, K.M.; Grimes, C.A.; Jung, J.-W.; Cho, C.-H.; Majima, T.; Hoffmann, M.R.; et al. CO2, Water, And Sunlight to Hydrocarbon Fuels: A Sustained Sunlight to Fuel (Joule-to-Joule) Photoconversion Efficiency of 1%. Energy Environ. Sci. 2019. [Google Scholar] [CrossRef]
  28. Kočí, K.; Reli, M.; Kozák, O.; Lacný, Z.; Plachá, D.; Praus, P.; Obalová, L. Influence of reactor geometry on the yield of CO2 photocatalytic reduction. Catal. Today 2011, 176, 212–214. [Google Scholar] [CrossRef]
  29. Nguyen, T.V.; Wu, J.C.S. Photoreduction of CO2 in an optical-fiber photoreactor: Effects of metals addition and catalyst carrier. Appl. Catal. A Gen. 2008, 335, 112–120. [Google Scholar] [CrossRef]
  30. Tahir, M.; Amin, N.A.S. Photocatalytic reduction of carbon dioxide with water vapors over montmorillonite modified TiO2 nanocomposites. Appl. Catal. B Environ. 2013, 142–143, 512–522. [Google Scholar] [CrossRef]
  31. Tahir, B.; Tahir, M.; Amin, N.S. Gold–indium modified TiO2 nanocatalysts for photocatalytic CO2 reduction with H2 as reductant in a monolith photoreactor. Appl. Surf. Sci. 2015, 338, 1–14. [Google Scholar] [CrossRef]
  32. Tahir, M.; Tahir, B.; Amin, N.S. Photocatalytic CO2 reduction by CH4 over montmorillonite modified TiO2 nanocomposites in a continuous monolith photoreactor. Mater. Res. Bull. 2015, 63, 13–23. [Google Scholar] [CrossRef]
  33. Wang, W.; Ku, Y. Photocatalytic degradation of gaseous benzene in air streams by using an optical fiber photoreactor. J. Photochem. Photobiol. A Chem. 2003, 159, 47–59. [Google Scholar] [CrossRef]
  34. Usubharatana, P.; McMartin, D.; Veawab, A.; Tontiwachwuthikul, P. Photocatalytic process for CO2 emission reduction from industrial flue gas streams. Ind. Eng. Chem. Res. 2006, 45, 2558–2568. [Google Scholar] [CrossRef]
  35. Liou, P.Y.; Chen, S.C.; Wu, J.C.S.; Liu, D.; MacKintosh, S.; Maroto-Valer, M.; Linforth, R. Photocatalytic CO2 reduction using an internally illuminated monolith photoreactor. Energy Environ. Sci. 2011, 4, 1487–1494. [Google Scholar] [CrossRef]
  36. Wu, J.C.S.; Wu, T.-H.; Chu, T.; Huang, H.; Tsai, D. Application of optical-fiber photoreactor for CO2 photocatalytic reduction. Top. Catal. 2008, 47, 131–136. [Google Scholar] [CrossRef]
  37. Wang, Z.Y.; Chou, H.C.; Wu, J.C.S.; Ping Tsai, D.; Mul, G.; Engineering, C. CO2 photoreduction using NiO/InTaO4 in optical-fiber reactor for renewable energy. Appl. Catal. A Gen. 2010, 380, 172–177. [Google Scholar] [CrossRef]
  38. Wang, W.; Ku, Y. The light transmission and distribution in an optical fiber coated with TiO2 particles. Chemosphere 2003, 50, 999–1006. [Google Scholar] [CrossRef]
  39. Ola, O.; Maroto-Valer, M.; Liu, D.; MacKintosh, S.; Lee, C.W.; Wu, J.C.S. Performance comparison of CO2 conversion in slurry and monolith photoreactors using Pd and Rh-TiO2 catalyst under ultraviolet irradiation. Appl. Catal. B Environ. 2012, 126, 172–179. [Google Scholar] [CrossRef]
  40. Camera-Roda, G.; Santarelli, F.; Martin, C.A. Design of photocatalytic reactors made easy by considering the photons as immaterial reactants. Sol. Energy 2005, 79, 343–352. [Google Scholar] [CrossRef]
  41. Paulino, P.N.; Salim, V.M.M.; Resende, N.S. Zn-Cu promoted TiO2 photocatalyst for CO2 reduction with H2O under UV light. Appl. Catal. B Environ. 2016, 185, 362–370. [Google Scholar] [CrossRef]
  42. Han, S.; Chen, Y.; Abanades, S.; Zhang, Z. Improving photoreduction of CO2 with water to CH4 in a novel concentrated solar reactor. J. Energy Chem. 2017, 26, 743–749. [Google Scholar] [CrossRef]
  43. Singh, V.; Beltran, I.J.C.; Ribot, J.C.; Nagpal, P. Photocatalysis deconstructed: Design of a new selective catalyst for artificial photosynthesis. Nano Lett. 2014, 14, 597–603. [Google Scholar] [CrossRef] [PubMed]
  44. Li, D.; Chen, Y.; Abanades, S.; Zhang, Z. Enhanced activity of TiO2 by concentrating light for photoreduction of CO2 with H2O to CH4. Catal. Commun. 2018, 113, 6–9. [Google Scholar] [CrossRef]
  45. Li, D.; Fang, X.; Liu, H.; Lu, H.; Zhang, Z. Photoreduction of CO2 to CH4 on g-C3N4: The effect of concentrating light and pretreatment. AIP Conf. Proc. 2018, 1971, 020006. [Google Scholar]
  46. Tan, L.-L.; Ong, W.-J.; Chai, S.-P.; Mohamed, A.R. Photocatalytic reduction of CO2 with H2O over graphene oxide-supported oxygen-rich TiO2 hybrid photocatalyst under visible light irradiation: Process and kinetic studies. Chem. Eng. J. 2017, 308, 248–255. [Google Scholar] [CrossRef]
  47. Guan, G.; Kida, T.; Yoshida, A. Reduction of carbon dioxide with water under concentrated sunlight using photocatalyst combined with Fe-based catalyst. Appl. Catal. B Environ. 2003, 41, 387–396. [Google Scholar] [CrossRef]
  48. Saladin, F.; Alxneit, I. Temperature dependence of the photochemical reduction of CO2 in the presence of H2O at the solid/gas interface of TiO2. J. Chem. Soc. Faraday Trans. 1997, 93, 4159–4163. [Google Scholar] [CrossRef]
  49. Tan, S.S.; Zou, L.; Hu, E. Kinetic modelling for photosynthesis of hydrogen and methane through catalytic reduction of carbon dioxide with water vapour. Catal. Today 2008, 131, 125–129. [Google Scholar] [CrossRef]
  50. Zhang, Q.H.; Han, W.D.; Hong, Y.J.; Yu, J.G. Photocatalytic reduction of CO2 with H2O on Pt-loaded TiO2 catalyst. Catal. Today 2009, 148, 335–340. [Google Scholar] [CrossRef]
  51. Yamashita, H.; Ikeue, K.; Takewaki, T.; Anpo, M. In situ XAFS studies on the effects of the hydrophobic-hydrophilic properties of Ti-Beta zeolites in the photocatalytic reduction of CO2 with H2O. Top. Catal. 2002, 18, 95–100. [Google Scholar] [CrossRef]
  52. Wu, J.C.S.; Lin, H.M.; Lai, C.L. Photo reduction of CO2 to methanol using optical-fiber photoreactor. Appl. Catal. A Gen. 2005, 296, 194–200. [Google Scholar] [CrossRef]
  53. Tahir, M.; Amin, N.A.S. Performance analysis of nanostructured NiO-In2O2/TiO2 catalyst for CO2 photoreduction with H2 in a monolith photoreactor. Chem. Eng. J. 2016, 285, 635–649. [Google Scholar] [CrossRef]
  54. Ali, S.; Razzaq, A.; In, S.-I. Development of graphene based photocatalysts for CO2 reduction to C1 chemicals: A brief overview. Catal. Today 2018, 335, 39–54. [Google Scholar] [CrossRef]
  55. Hiragond, C.; Ali, S.; Sorcar, S.; In, S.-I. Hierarchical Nanostructured Photocatalysts for CO2 Photoreduction. Catalysts 2019, 9, 370. [Google Scholar] [CrossRef]
  56. In, S.-I.; Vaughn, D.D.; Schaak, R.E. Hybrid CuO-TiO2−xNx Hollow Nanocubes for Photocatalytic Conversion of CO2 into Methane under Solar Irradiation. Angew. Chemie Int. Ed. 2012, 51, 3915–3918. [Google Scholar] [CrossRef]
  57. In, S.-I.; Orlov, A.; García, F.; Tikhov, M.; Wright, D.S.; Lambert, R.M. Efficient visible light-active N-doped TiO2 photocatalysts by a reproducible and controllable synthetic route. Chem. Commun. 2006, 4236–4238. [Google Scholar] [CrossRef] [PubMed]
  58. In, S.-I.; Orlov, A.; Berg, R.; García, F.; Pedrosa-Jimenez, S.; Tikhov, M.S.; Wright, D.S.; Lambert, R.M. Effective visible light-activated B-doped and B, N-codoped TiO2 photocatalysts. J. Am. Chem. Soc. 2007, 129, 13790–13791. [Google Scholar] [CrossRef]
  59. Cortes, M.; Hamilton, J.W.J.; Sharma, P.K.; Brown, A.; Nolan, M.; Gray, K.A.; Byrne, J.A. Formal quantum efficiencies for the photocatalytic reduction of CO2 in a gas phase batch reactor. Catal. Today 2019, 326, 75–81. [Google Scholar] [CrossRef]
  60. Tahir, M.; Amin, N.S. Indium-doped TiO2 nanoparticles for photocatalytic CO2 reduction with H2O vapors to CH4. Appl. Catal. B Environ. 2015, 162, 98–109. [Google Scholar] [CrossRef]
  61. Wang, M.; Han, Q.; Li, L.; Tang, L.; Li, H.; Zhou, Y.; Zou, Z. Construction of an all-solid-state artificial Z-scheme system consisting of Bi2WO6/Au/CdS nanostructure for photocatalytic CO2 reduction into renewable hydrocarbon fuel. Nanotechnology 2017, 28, 274002. [Google Scholar] [CrossRef] [PubMed]
  62. 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]
  63. Kozlova, E.A.; Lyulyukin, M.N.; Markovskaya, D.V.; Selishchev, D.S.; Cherepanova, S.V.; Kozlov, D.V. Synthesis of Cd1−xZnxS photocatalysts for gas-phase CO2 reduction under visible light. Photochem. Photobiol. Sci. 2019, 18, 871–877. [Google Scholar] [CrossRef] [PubMed]
  64. Yan, S.; Ouyang, S.; Xu, H.; Zhao, M.; Zhang, X.; Ye, J. Co-ZIF-9/TiO2 nanostructure for superior CO2 photoreduction activity. J. Mater. Chem. A 2016, 4, 15126–15133. [Google Scholar] [CrossRef]
  65. Li, X.; Zhuang, Z.; Li, W.; Pan, H. Photocatalytic reduction of CO2 over noble metal-loaded and nitrogen-doped mesoporous TiO2. Appl. Catal. A Gen. 2012, 429, 31–38. [Google Scholar] [CrossRef]
  66. Cao, S.-W.; Liu, X.-F.; Yuan, Y.-P.; Zhang, Z.-Y.; Liao, Y.-S.; Fang, J.; Loo, S.C.J.; Sum, T.C.; Xue, C. Solar-to-fuels conversion over In2O3/g-C3N4 hybrid photocatalysts. Appl. Catal. B Environ. 2014, 147, 940–946. [Google Scholar] [CrossRef]
  67. Hong, J.; Zhang, W.; Wang, Y.; Zhou, T.; Xu, R. Photocatalytic Reduction of Carbon Dioxide over Self-Assembled Carbon Nitride and Layered Double Hydroxide: The Role of Carbon Dioxide Enrichment. ChemCatChem 2014, 6, 2315–2321. [Google Scholar] [CrossRef]
  68. Li, Y.; Wang, C.; Song, M.; Li, D.; Zhang, X.; Liu, Y. TiO2−x/CoOx photocatalyst sparkles in photothermocatalytic reduction of CO2 with H2O steam. Appl. Catal. B Environ. 2019, 243, 760–770. [Google Scholar] [CrossRef]
  69. Tahir, M.; Tahir, B.; Amin, N.A.S. Synergistic effect in plasmonic Au/Ag alloy NPs co-coated TiO2 NWs toward visible-light enhanced CO2 photoreduction to fuels. Appl. Catal. B Environ. 2017, 204, 548–560. [Google Scholar] [CrossRef]
  70. Pan, B.; Luo, S.; Su, W.; Wang, X. Photocatalytic CO2 reduction with H2O over LaPO4 nanorods deposited with Pt cocatalyst. Appl. Catal. B Environ. 2015, 168, 458–464. [Google Scholar] [CrossRef]
  71. Zhang, N.; Ouyang, S.; Li, P.; Zhang, Y.; Xi, G.; Kako, T.; Ye, J. Ion-exchange synthesis of a micro/mesoporous Zn2GeO4 photocatalyst at room temperature for photoreduction of CO2. Chem. Commun. 2011, 47, 2041–2043. [Google Scholar] [CrossRef] [PubMed]
  72. Jiao, X.; Chen, Z.; Li, X.; Sun, Y.; Gao, S.; Yan, W.; Wang, C.; Zhang, Q.; Lin, Y.; Luo, Y. Defect-mediated electron–hole separation in one-unit-cell ZnIn2S4 layers for boosted solar-driven CO2 reduction. J. Am. Chem. Soc. 2017, 139, 7586–7594. [Google Scholar] [CrossRef] [PubMed]
  73. An, X.; Li, K.; Tang, J. Cu2O/reduced graphene oxide composites for the photocatalytic conversion of CO2. ChemSusChem 2014, 7, 1086–1093. [Google Scholar] [CrossRef] [PubMed]
  74. Chong, R.; Su, C.; Du, Y.; Fan, Y.; Ling, Z.; Chang, Z.; Li, D. Insights into the role of MgAl layered double oxides interlayer in Pt/TiO2 toward photocatalytic CO2 reduction. J. Catal. 2018, 363, 92–101. [Google Scholar] [CrossRef]
  75. Yang, Y.; Wu, J.; Xiao, T.; Tang, Z.; Shen, J.; Li, H.; Zhou, Y.; Zou, Z. Urchin-like Hierarchical CoZnAl-LDH/RGO/g-C3N4 Hybrid as a Z-Scheme Photocatalyst for Efficient and Selective CO2 Reduction. Appl. Catal. B Environ. 2019, 255, 117771. [Google Scholar] [CrossRef]
  76. Shi, W.; Guo, X.; Cui, C.; Jiang, K.; Li, Z.; Qu, L.; Wang, J.-C. Controllable synthesis of Cu2O decorated WO3 nanosheets with dominant (001) facets for photocatalytic CO2 reduction under visible-light irradiation. Appl. Catal. B Environ. 2019, 243, 236–242. [Google Scholar] [CrossRef]
  77. Fang, B.; Bonakdarpour, A.; Reilly, K.; Xing, Y.; Taghipour, F.; Wilkinson, D.P. Large-scale synthesis of TiO2 microspheres with hierarchical nanostructure for highly efficient photodriven reduction of CO2 to CH4. ACS Appl. Mater. Interfaces 2014, 6, 15488–15498. [Google Scholar] [CrossRef] [PubMed]
  78. Yu, J.; Jin, J.; Cheng, B.; Jaroniec, M. A noble metal-free reduced graphene oxide–CdS nanorod composite for the enhanced visible-light photocatalytic reduction of CO2 to solar fuel. J. Mater. Chem. A 2014, 2, 3407–3416. [Google Scholar] [CrossRef]
  79. Wang, S.; Xu, M.; Peng, T.; Zhang, C.; Li, T.; Hussain, I.; Wang, J.; Tan, B. Porous hypercrosslinked polymer-TiO2-graphene composite photocatalysts for visible-light-driven CO2 conversion. Nat. Commun. 2019, 10, 676. [Google Scholar] [CrossRef]
  80. Li, Q.; Lin, F.; Liu, F.; Wang, X. A CO2 photo-reduction heterogeneous cobalt-based cocatalyst by in-situ electrostatic adsorption deposition. Chem. Commun. 2019, 55, 3903–3906. [Google Scholar] [CrossRef]
  81. Fang, B.; Xing, Y.; Bonakdarpour, A.; Zhang, S.; Wilkinson, D.P. Hierarchical CuO–TiO2 hollow microspheres for highly efficient photodriven reduction of CO2 to CH4. ACS Sustain. Chem. Eng. 2015, 3, 2381–2388. [Google Scholar] [CrossRef]
  82. Long, R.; Li, Y.; Liu, Y.; Chen, S.; Zheng, X.; Gao, C.; He, C.; Chen, N.; Qi, Z.; Song, L. Isolation of Cu atoms in Pd lattice: Forming highly selective sites for photocatalytic conversion of CO2 to CH4. J. Am. Chem. Soc. 2017, 139, 4486–4492. [Google Scholar] [CrossRef] [PubMed]
  83. Tahir, M. Hierarchical 3D VO2/ZnV2O4 microspheres as an excellent visible light photocatalyst for CO2 reduction to solar fuels. Appl. Surf. Sci. 2019, 467, 1170–1180. [Google Scholar] [CrossRef]
  84. Xu, F.; Meng, K.; Cheng, B.; Yu, J.; Ho, W. Enhanced Photocatalytic Activity and Selectivity for CO2 Reduction over a TiO2 Nanofibre Mat Using Ag and MgO as Bi-Cocatalyst. ChemCatChem 2019, 11, 465–472. [Google Scholar] [CrossRef]
  85. Ye, L.; Wang, H.; Jin, X.; Su, Y.; Wang, D.; Xie, H.; Liu, X.; Liu, X. Synthesis of olive-green few-layered BiOI for efficient photoreduction of CO2 into solar fuels under visible/near-infrared light. Sol. Energy Mater. Sol. Cells 2016, 144, 732–739. [Google Scholar] [CrossRef]
  86. Jin, J.; Yu, J.; Guo, D.; Cui, C.; Ho, W. A Hierarchical Z-Scheme CdS–WO3 Photocatalyst with Enhanced CO2 Reduction Activity. Small 2015, 11, 5262–5271. [Google Scholar] [CrossRef] [PubMed]
  87. Hou, T.; Luo, N.; Cui, Y.-T.; Lu, J.; Li, L.; MacArthur, K.E.; Heggen, M.; Chen, R.; Fan, F.; Tian, W. Selective reduction of CO2 to CO under visible light by controlling coordination structures of CeOx-S/ZnIn2S4 hybrid catalysts. Appl. Catal. B Environ. 2019, 245, 262–270. [Google Scholar] [CrossRef]
  88. Lin, L.-Y.; Nie, Y.; Kavadiya, S.; Soundappan, T.; Biswas, P. N-doped reduced graphene oxide promoted nano TiO2 as a bifunctional adsorbent/photocatalyst for CO2 photoreduction: Effect of N species. Chem. Eng. J. 2017, 316, 449–460. [Google Scholar] [CrossRef]
  89. Reñones, P.; Moya, A.; Fresno, F.; Collado, L.; Vilatela, J.J.; Víctor, A. Hierarchical TiO2 nanofibres as photocatalyst for CO2 reduction: Influence of morphology and phase composition on catalytic activity. J. CO2 Util. 2016, 15, 24–31. [Google Scholar] [CrossRef]
  90. Shown, I.; Hsu, H.-C.; Chang, Y.-C.; Lin, C.-H.; Roy, P.K.; Ganguly, A.; Wang, C.-H.; Chang, J.-K.; Wu, C.-I.; Chen, L.-C. Highly efficient visible light photocatalytic reduction of CO2 to hydrocarbon fuels by Cu-nanoparticle decorated graphene oxide. Nano Lett. 2014, 14, 6097–6103. [Google Scholar] [CrossRef]
  91. Xiong, Z.; Luo, Y.; Zhao, Y.; Zhang, J.; Zheng, C.; Wu, J.C.S. Synthesis, characterization and enhanced photocatalytic CO2 reduction activity of graphene supported TiO2 nanocrystals with coexposed {001} and {101} facets. Phys. Chem. Chem. Phys. 2016, 18, 13186–13195. [Google Scholar] [CrossRef] [PubMed]
  92. Kong, X.Y.; Lee, W.Q.; Mohamed, A.R.; Chai, S.-P. Effective Steering of Charge Flow through Synergistic Inducing Oxygen Vacancy Defects and pn Heterojunctions in 2D/2D Surface-Engineered Bi2WO6−x/BiOI Cascade: Towards Superior Photocatalytic CO2 Reduction Activity. Chem. Eng. J. 2019, 372, 1183–1193. [Google Scholar] [CrossRef]
  93. Xiong, Z.; Wang, H.; Xu, N.; Li, H.; Fang, B.; Zhao, Y.; Zhang, J.; Zheng, C. Photocatalytic reduction of CO2 on Pt2+–Pt0/TiO2 nanoparticles under UV/Vis light irradiation: A combination of Pt2+ doping and Pt nanoparticles deposition. Int. J. Hydrogen Energy 2015, 40, 10049–10062. [Google Scholar] [CrossRef]
  94. Zhao, Y.; Wei, Y.; Wu, X.; Zheng, H.; Zhao, Z.; Liu, J.; Li, J. Graphene-wrapped Pt/TiO2 photocatalysts with enhanced photogenerated charges separation and reactant adsorption for high selective photoreduction of CO2 to CH4. Appl. Catal. B Environ. 2018, 226, 360–372. [Google Scholar] [CrossRef]
  95. Mateo, D.; Albero, J.; García, H. Graphene supported NiO/Ni nanoparticles as efficient photocatalyst for gas phase CO2 reduction with hydrogen. Appl. Catal. B Environ. 2018, 224, 563–571. [Google Scholar] [CrossRef]
  96. Gangopadhyay, S.; Wang, W.-N.; Biswas, P.; Mukherjee, S.; Ramalingam, B.; An, W.-J.; Niedzwiedzki, D.M. Size and Structure Matter: Enhanced CO2 Photoreduction Efficiency by Size-Resolved Ultrafine Pt Nanoparticles on TiO2 Single Crystals. J. Am. Chem. Soc. 2012, 134, 11276–11281. [Google Scholar]
Figure 1. Fourier Transform Infrared Spectroscopy FTIR spectra of Cu(I)/TiO2 preloaded with 13CO2 after 80-min illumination. (a) fresh Cu(I)/TiO2 (synthesized with PEG), (b) Cu(I)/TiO2 cleaned by illumination in humid air for 14 h, and (c) reference Cu(I)/TiO2 (synthesized without polyethylene glycol (PEG)), reproduced with permission from reference [4]. Copyright American Chemical Society, 2010.
Figure 1. Fourier Transform Infrared Spectroscopy FTIR spectra of Cu(I)/TiO2 preloaded with 13CO2 after 80-min illumination. (a) fresh Cu(I)/TiO2 (synthesized with PEG), (b) Cu(I)/TiO2 cleaned by illumination in humid air for 14 h, and (c) reference Cu(I)/TiO2 (synthesized without polyethylene glycol (PEG)), reproduced with permission from reference [4]. Copyright American Chemical Society, 2010.
Catalysts 09 00727 g001
Figure 2. (ac) production of CH4, C2H6, and C2H4 over Ti-SBA-15 for four cycles, and (d) blank test (without catalyst). After every cycle reactor was evacuated with He/H2O, reproduced with permission from reference [5]. Copyright Elsevier, 2011.
Figure 2. (ac) production of CH4, C2H6, and C2H4 over Ti-SBA-15 for four cycles, and (d) blank test (without catalyst). After every cycle reactor was evacuated with He/H2O, reproduced with permission from reference [5]. Copyright Elsevier, 2011.
Catalysts 09 00727 g002
Figure 3. CO production for ZIF-8/TiO2 under constant pressure (CP, 5 bar) and pressure swing (PS, 5–3 bar) at flow rates of 50 and 100 mL min−1, reproduced with permission from reference [19]. Copyright Elsevier, 2017.
Figure 3. CO production for ZIF-8/TiO2 under constant pressure (CP, 5 bar) and pressure swing (PS, 5–3 bar) at flow rates of 50 and 100 mL min−1, reproduced with permission from reference [19]. Copyright Elsevier, 2017.
Catalysts 09 00727 g003
Figure 4. Illustration of twin reactor setup, reproduced with permission from reference [22]. Copyright Elsevier, 2018.
Figure 4. Illustration of twin reactor setup, reproduced with permission from reference [22]. Copyright Elsevier, 2018.
Catalysts 09 00727 g004
Figure 5. Setup for continuous gas phase photocatalytic CO2 reduction, reproduced with permission from reference [11]. Copyright Elsevier, 2017.
Figure 5. Setup for continuous gas phase photocatalytic CO2 reduction, reproduced with permission from reference [11]. Copyright Elsevier, 2017.
Catalysts 09 00727 g005
Figure 6. Reactor diagram and its blue prints, (a) Complete reactor diagram, (b) Lid dimensions, and (c) Reactor body dimensions.
Figure 6. Reactor diagram and its blue prints, (a) Complete reactor diagram, (b) Lid dimensions, and (c) Reactor body dimensions.
Catalysts 09 00727 g006
Figure 7. Illustration for comparison of TiO2 coated on monolith and cell type support, reproduced with permission from reference [6]. Copyright Elsevier, 2013.
Figure 7. Illustration for comparison of TiO2 coated on monolith and cell type support, reproduced with permission from reference [6]. Copyright Elsevier, 2013.
Catalysts 09 00727 g007
Figure 8. Illustration of monolith fiber optic reactor, reproduced from reference [35] with permission from The Royal Society of Chemistry.
Figure 8. Illustration of monolith fiber optic reactor, reproduced from reference [35] with permission from The Royal Society of Chemistry.
Catalysts 09 00727 g008
Figure 9. Temperature dependent production of CO, reproduced with permission from reference [6]. Copyright Elsevier, 2013.
Figure 9. Temperature dependent production of CO, reproduced with permission from reference [6]. Copyright Elsevier, 2013.
Catalysts 09 00727 g009
Figure 10. Effect of H2O/CO2 ratio over CH4 yield, reproduced with permission from reference [30]. Copyright Elsevier, 2013.
Figure 10. Effect of H2O/CO2 ratio over CH4 yield, reproduced with permission from reference [30]. Copyright Elsevier, 2013.
Catalysts 09 00727 g010
Figure 11. Performance evaluation of cell type and monolith batch/continuous photoreactor for CO2 photoreduction, reproduced with permission from reference [53]. Copyright Elsevier, 2015.
Figure 11. Performance evaluation of cell type and monolith batch/continuous photoreactor for CO2 photoreduction, reproduced with permission from reference [53]. Copyright Elsevier, 2015.
Catalysts 09 00727 g011
Figure 12. Effect of different light intensities at (a) CH4 yield and (b) apparent quantum yield(AQY) at variable light intensities, reproduced with permission from reference [46]. Copyright Elsevier, 2016.
Figure 12. Effect of different light intensities at (a) CH4 yield and (b) apparent quantum yield(AQY) at variable light intensities, reproduced with permission from reference [46]. Copyright Elsevier, 2016.
Catalysts 09 00727 g012
Table 1. Summary of operating parameters and yields reported by In et al. for photocatalytic CO2 reduction.
Table 1. Summary of operating parameters and yields reported by In et al. for photocatalytic CO2 reduction.
PhotocatalystSynthesis MethodPre-Treatment of ReactorLight SourceReducing AgentReaction ParametersReactorMain ProductAQY (%)
Degussa P25
standard titania
[23]
store-boughtpurged with high purity CO2 gas, at least five timesUVP, UVGL-58 lamp with λ = 365 nm; 1200 μW cm−2H2O50 mg catalyst on a 30 mm diam. glass disk; 15.4 cm3 reactor; CO2 flowrate @ 10 cm3 min−1; 500 μL sample gas extracted; ambient temperature and pressure;
1 h irradiation
Batch reactorCH4 @ 0.021
µmol g−1 h−1
0.0021
CZTS−TiO2
hybrid mesoporous
[24]
hot injection and annealingpurged with CO2 gas (1000 ppm in He) and vacuum100 W Xe solar simulator with an AM 1.5 filter;
100 mW cm−2
H2O50 mg catalyst on a 30 mm diam. glass disk; 15.4 cm3 reactor; CO2 flowrate @ 10 cm3 min−1; 500 μL sample gas extracted; ambient temperature and pressure;
1 h irradiation
Batch reactorCH4 @ 118.75
ppm g−1 h−1
0.0057
CZTS-ZnO
nanoparticles
[14]
hydrothermal treatmentthree times purged with CO2 gas (1000 ppm in He) and vacuum100 W Xe solar simulator with an AM 1.5 filter;
100 mW cm−2
H2O50 mg catalyst on a 30 mm diam. glass disk; 15.4 cm3 reactor; CO2 flowrate @ 10 cm3 min−1; 500 μL sample gas extracted every 1 h; ambient temperature and pressure;
1 h irradiation
Batch reactorCH4 @ 0.0954
µmol g−1 h−1
0.0128
CuxO−TiO2
mesoporous p-type/n-type heterojunction material
[25]
thermal decomposition then calcinationpurged with CO2 gas (1000 ppm in He) and vacuum100 W Xe solar simulator with an AM 1.5 filter;
100 mW cm−2
H2O50 mg catalyst on a 30 mm diam. glass disk; 15.4 cm3 reactor; CO2 flowrate @ 10 cm3 min−1; 500 μL sample gas extracted; ambient temperature and pressure;
1 h irradiation
Batch reactorCH4 @ 221.63
ppm g−1 h−1
0.0177
Pt-x-RT
nanoparticles
[15]
magnesio-thermic reductionfive times purged with CO2 gas (1000 ppm in He) and vacuum100 W Xe solar simulator with an AM 1.5 filter;
100 mW cm−2
H2O70 mg catalyst on a 30 mm diam. glass disk; 15.4 cm3 reactor; CO2 flowrate @ 10 cm3 min−1; 500 μL sample gas extracted; ambient temperature and pressure;
1 h irradiation
Batch reactorCH4 @ 1.13
µmol g−1 h−1
0.1234
C,N-TNT06
nanotubes
[26]
alkaline hydrothermal technique then calcinationpurged with CO2 gas (1000 ppm in He) and vacuum100 W Xe solar simulator with an AM 1.5 filter;
100 mW cm−2
H2O100 mg catalyst on a 30 mm diam. glass disk; 15.4 cm3 reactor; CO2 flowrate @ 10 cm3 min−1; 500 μL sample gas extracted; ambient temperature and pressure;
1 h irradiation
Batch reactorCH4 @ 9.75
µmol g−1 h−1
1.0532
Pt-XG/RBT
nanoparticles
[8]
facile vacuum treatment and photodeposition1 h purging with moist CO2 gas @ 40 mL min−1100 W Xe solar simulator with an AM 1.5 filter;
100 mW cm−2
H2O40 mg catalyst on a 4.9 cm2 porous disk; 26.57 cm3 reactor; CO2 flowrate @ 1 mL min−1; sample gas analyzed every 30 min; ambient temperature and pressure;
7 h irradiation
Continuous flow reactorCH4 @ 37.0
µmol g−1 h−1
(AQYCH4 = 5.248)
C2H6 @ 11.0
µmol g−1 h−1
(AQYC2H6 = 2.73)
7.978
Pt-BT-X
nanoparticles
[11]
facile low-temperature synthesis, annealing and photodeposition1 h purging with moist CO2 gas @ 40 mL min−1100 W Xe solar simulator with an AM 1.5 filter;
100 mW cm−2
H2O40 mg catalyst on a 4.9 cm2 porous disk; 26.57 cm3 reactor; CO2 flowrate @ 1 mL min−1; sample gas analyzed every 30 min; ambient temperature and pressure;
6 h irradiation
Continuous flow reactorCH4 @ 80.35
µmol g−1 h−1
12.357
Cux%–Pty%–BT
nanoparticles
[27]
facile low-temperature synthesis, annealing and photodeposition1 h purging with moist CO2 gas @ 40 cc min−1100 W Xe solar simulator with an AM 1.5 filter;
100 mW cm−2
H2O40 mg catalyst on a 2.5 cm diam. porous disk; 26.57 cm3 reactor; CO2 flowrate @ 1 mL min−1; sample gas analyzed every 30 min; ambient temperature and pressure;
6 h irradiation
Continuous flow reactorCH4 @ 3.0
mmol g−1
(AQYCH4 = 79.14)
C2H6 @ 0.15
mmol g−1
(AQYC2H6 = 6.92)
86
Table 2. Summary of operating parameters and yields used by other research works for photocatalytic CO2 reduction using Batch reactor.
Table 2. Summary of operating parameters and yields used by other research works for photocatalytic CO2 reduction using Batch reactor.
PhotocatalystSynthesis MethodPre-Treatment of ReactorLight SourceReducing AgentReaction ParametersReactorMain ProductAQY (%)
Bi2WO6/Au/CdS
Z-scheme
[61]
bath deposition methodvacuum-treated several times, and then filled with high purity CO2 gas300 W Xe lamp
(λ > 400 nm)
0.4 mL of DI water100 mg catalyst; 230 mL reactor; 1 mL sample gas;
ambient pressure;
8 h irradiation
Batch reactorCH4 @ ~0.75
µmol g−1 h−1
0.012 b
rGO-CuO
hybrid structure
[62]
covalent graftingpurged with nitrogen gas for 15 min then purged with CO2 for 30 min under continuous stirring20 W white cold LED flood light
(200 < λ < 700 nm); 85 W m−2
mixture of DMF (45 mL) and H2O (5 mL)100 mg catalyst; 100 mL reactor; 20 μL total sample gas analyzed;
24 h irradiation
Batch reactorCH3OH @ 1282
µmol g−1
0.013 b,c
Cd1−xZnxS
solid solution
[63]
two-step self-templated synthesispurged with argon for 1 h, then 2 mL of deionized water was injected, and purged with ultra-pure CO2 for 30 min100 W LED plate with collimating lens; visible light (λ = 450 nm); 285 mW cm−22 mL of DI water45 mg catalyst; 130 cm3 reactor; 250 μL sample gas extracted every 1 h;
1 atm; 25 °C;
5 h irradiation
Batch reactorCO @ 2.90
µmol g−1 h−1
(AQYCO = 0.016)
CH4 @ 0.22 µmol g−1 h−1
(AQYCH4 = 0.005)
0.02 a
Co-ZIF-9/TiO2
nanostructure
[64]
in situ growth methodpurged with high-purity CO2 gas300 W Xe lamp (200 < λ < 900 nm); 494 mW cm−23 mL DI water50 mg catalyst; 390 mL reactor; 0.5 mL sample gas extracted; 70 kPa;
10 h irradiation
Batch reactor with gas circulation systemCO @ 17.58
µmol g−1 h−1
0.053 a
Pt/TiO2
mesoporous structure
[65]
soft-template methodpurged with high purity CO2 bubbled through DI water for more than 1 h350 W Xe-lamp with 420 nm cutoff filter;
UV light @
34.8 mW cm−2
H2O100 mg catalyst; 159 mL tubular reactor (length: 28 cm, Ø: 3 cm); 60 ± 2 °C; 2 h irradiationBatch reactorCH4 @ 5.7
µmol g−1
0.064 a
In2O3–C3N4
hybrid structure
[66]
simple solvothermal methodpurged with high-purity CO2 gas500 W
Xenon lamp;
1200 mW cm−2
0.1 mL ultrapure H2O20 mg catalyst; 90 mL reactor; 4 h irradiationBatch reactorCH4 @ 159.2 ppm0.082 a
Pd/(10 wt.% LDH/C3N4)
hybrid structure
[67]
electrostatic interactionintroduction of 200 torr CO2 into the system500 W Hg (Xe) lamp without filter100 mL H2O200 mg catalyst; 200 μL sample gas extracted; AQY @ λ = 420 nm; 200 torr; 72 h irradiationBatch reactorCH4 @ 6.5 µmol0.093 b
In/TiO2-monolith
[6]
sol–gel single step methodcontinuous passing of CO2, He and H2O mixture through the reactor for about 1 h200 W Hg lamp for UV irradiations
(λ < 252 nm);
150 mW cm−2
H2O50 mg catalyst; 150 cm3 reactor; 1000 μL sample gas extracted; PCO2 = 0.20 bar; PH2O = 0.074 bar;
10 h irradiation
Batch reactorCO @ 962
µmol g−1 h−1
0.10 b
TiO2−x/CoOxhybrid structure
[68]
(own method)blown with CO2 for 20 min150 W UV lamp;
20 mW cm−2
2 mL of DI water50 mg catalyst;
100 mL reactor; 1.01 bar; room temperature;
4 h irradiation
Batch reactorCO @ 1.247
µmol g−1 h−1
(AQYCO = 0.0817)
CH4 @ 0.0903
µmol g−1 h−1
(AQYCH4 = 0.0237)
0.105 a
Ag-Au/TiO2
nanowires
[69]
facile hydrothermal synthesiscompressed CO2 and H2 were continuously passed through the reactor35 W HID Xe lamp;
20 mW cm−2
H210 mg catalyst; 108 cm3 reactor; 0.20 bar;
4 h irradiation
Batch reactorCO @ 1813
µmol g−1 h−1
0.1108 b
LaPO4–Pt
nanorods
[70]
hydrothermal method and photo depositionreactor was evacuated and filled with CO2 for 1 h with stirring125 W high-pressure Hg lamp
(λ < 365 nm)
70 mL H2O50 mg catalyst; 200 mL reactor; 1 atm; 20 °C;
4 h irradiation
Batch reactorCH4 @ 0.62 µmol g−10.15 b
Zn2GeO4
micro/mesoporous
[71]
simple ion exchangevacuum-pumped and washed with high purity CO2 gas300 W Xe arc lamp
(λ = 251 ± 16 nm)
0.5 mL DI water200 mg catalyst; 360 mL reactor; 0.5 mL sample gas extracted; ambient pressure; 12 h irradiationBatch reactorCH4 @ 9.5 ppm g−1 h−10.20 b
ZnIn2S4
one-unit-cell atomic layers
[72]
(own method)vacuum-treated three times, then pumped with high-purity CO2PLS-SXE300/
300 UV Xe lamp;
100 mW cm−2
2 mL DI water100 mg catalyst;
atmospheric pressure; 298 ± 0.2 K; 1 h irradiation
Batch reactorCO @ 33.2
µmol g−1 h−1
0.23 b
Cu2O/x% RGO
composites
[73]
microwave-assisted hydrothermal reactionCO2 purged150 W Xe lamp;
540 µW cm−2
3 mL DI water500 mg catalyst; 120 mL reactor; sample gas extracted every 30 min;
20 h irradiation
Batch reactorCO @ 50
ppm g−1 h−1
0.34 b
Pt/MgAl-LDO/TiO2
hybrid structure
[74]
in-situ deposition, calcination and photo depositiondegassed for 30 min, and then bubbled with CO2 till the pressure reaches 1 atm300 W Xe lamp;
1.1 mW cm2
H2O20 mg catalyst; AQY @
λ = 365 nm; 1 atm; 20 °C; 8 h irradiation
Batch reactorCH4 @
0.11 µmol
0.35 b,c
LDH/RGO/CN
hybrid structure
[75]
hydrothermal synthesis and in situ loadingvacuum-treated several times, and then flowed with high purity CO2 gas300 W
Xe arc lamp;
1.8 mW cm−2
4 mL DI water50 mg catalyst; 420 mL reactor; 1 mL sample gas extracted; AQY @ λ = 385 nm; ambient pressure; 5 h irradiationBatch reactorCO @ 10.11
µmol g−1 h−1
0.45 b
Cu2O/WO3
nanosheets
[76]
modified methodvacuum treated, and then purged several times with high purity CO2 gas300 W Xenon arc lamp with a UV cutoff filter (λ > 400 nm)H2O85 mg catalyst;
18 h irradiation
Batch reactorCO @ 0.56
µmol g−1 h−1
0.503 b
TiO2
microsphere
[77]
sol−gel approachintroduction of pressurized CO2
@ (50 psi)
40 W Hg UV lamp
(λ = 254 nm);
20 mW cm−2
100 μL H2O200 mg catalyst; 39 mm diameter and 9 mm depth reactor; 10 mL sample gas extracted; 50 psi;
24 h irradiation
Batch reactorCO @ 0.56
µmol g−1 h−1
(AQYCO = 0.204)
CH4 @ 0.94
µmol g−1 h−1
(AQYCH4 = 0.34)
0.54 b,c
RGO-CdS
nanorod composites
[78]
microwave hydrothermal routedegassed with nitrogen for 30 min300 W Xe arc lamp with a UV-cutoff filter (λ ≥ 420 nm);
150 mW cm−2
10 mL distilled water100 mg catalyst; 200 mL reactor; 1 mL sample gas extracted every 1 h; atmospheric pressure and ambient temperature;
3 h irradiation
Batch reactorCH4 @ 2.51
µmol g−1 h−1
0.80 b
HCP-TiO2-FG
composite
[79]
in situ growth-300 W Xe lamp
(λ ≥ 420 nm); 433 mW cm−2
H2O20 mg catalyst; standard atmospheric pressure;
5 h irradiation
Batch reactorCH4 @ 27.62
µmol g−1 h−1
(AQYCH4 = 1.14)
CO @ 21.63
µmol g−1 h−1
(AQYCO = 0.2227)
1.36 a
Co/Palheterostructure
[80]
in situ electrostatic adsorption deposition processfilled with high purity CO2 gas300 W Xe lamp5 mL acetonitrile/H2O (4:1)9 mg photosensitizer + 1 mg co-catalyst + 1 mL TEOA; 80 mL reactor; AQY @ λ = 420 nm; 1 atm; 25 °C; 6 h irradiationBatch reactorCO @ ~86 µmol1.38 b
CuO−TiO2
hollow microspheres
[81]
one-pot template-free synthesisintroduction of pressurized CO2
@ (50 psi)
40 W Hg UV lamp(λ = 254 nm);
20 mW cm−2
200 μL H2O10 mg catalyst; reactor diameter of 39 mm and a depth of 9 mm; 50 psi;
24 h irradiation
Batch reactorCO @ 14.5
µmol g−1 h−1
(AQYCO = 1.285)
CH4 @ 2.1
µmol g−1 h−1
(AQYCH4 = 0.747)
2.03 b
Pt-TiO2
spheres
[77]
microwave-assisted solvothermal methodintroduction of pressurized CO2
@ (50 psi)
40 W Hg UV lamp
(λ = 254 nm);
20 mW cm−2
100 μL H2O200 mg catalyst;
39 mm diameter and 9 mm depth reactor; 10 mL sample gas extracted; 50 psi; 24 h irradiation
Batch reactorCO @ 18.9
µmol g−1 h−1
(AQYCO = 1.632)
CH4 @ 3.6
µmol g−1 h−1
(AQYCH4 = 1.315)
2.95 b,c
PdxCu1-TiO2
hybrid structures
[82]
in situ growthfilled with
0.2 MPa CO2 for 60 min
300 W Xe lamp
(λ < 400 nm);
2 mW cm−2
H2O5 mg of TiO2 + 0.01 mmol of metal atoms for catalyst; 100 mL reactor; 0.2 MPa;
2 h irradiation
Batch reactorCH4 @ 19.6
µmol g−1 h−1
12.53 a
In/TiO2
nanoparticles
[60]
sol–gel single step methodpurged with CO2 and He for an hour500 W mercury flash lamp
(λ = 365 nm);
40 mW cm−2
H2O0.25 mg catalyst; 106 cm3 reactor; 1000 μL sample gas extracted; 0.20 bars, 373 K; 8 h irradiationBatch reactorCH4 @ 244
µmol g−1 h−1
(AQYCH4 = 42.39)
CO @ 81
µmol g−1 h−1
(AQYCO = 3.52)
45.91 a
ZnV2O4
microspheres
[83]
one-step hydrothermal processpurged with CO2 gas carrying H2O for 30 min35 W HID
Xe lamp;
20 mW cm−2
H2O100 mg catalyst; CO2 flowrate @ 20 mL min−1; 108 cm3 reactor; 0.20 bar; 100 °C; 4 h irradiationBatch reactorCO @ 485
µmol g−1 h−1
(AQYCO = 31.92)
CH3OH @ 100
µmol g−1 h−1
(AQYCH3OH = 19.75)
51.67 a
NiO/InTaO4
monolith coated structure
[35]
impregnation method and sol-gel methodpurged overnight using a flow of He then switched to pure CO2 with saturated water vapor for 1 h300 W Xe arc lamp with AM 1.5 filter;
100 mW cm−2
H2O88.7 mg catalyst; 216 cm3 reactor; 1 bar; 70 °C;
2 h irradiation
-CH3OH @ 0.16
µmol g−1 h−1
(AQYCH3OH = 0.012)
CH3CHO @ 0.3
µmol g−1 h−1
(AQYCH3CHO = 0.058)
0.07 b
MAT
nanofibers
[84]
(own method)blown with nitrogen for 30 min300 W simulated solar Xe arc lampH2O200 mL reactor; 1 mL sample gas extracted every 1 h; atmospheric pressure and ambient temperature;
3 h irradiation
-CH4 @ 0.86
µmol g−1 h−1
0.091 b
BiOI
few-layered nanosheets
[85]
(own method)thoroughly vacuum-treated300 W high pressure Xe lamp5 mL H2SO4 & 1.712 g NaHCO3150 mg catalyst; 500 mL reactor; 0.15 mL sample gas extracted; 20 °C
AQY @ λ = 420 nm;
4 h irradiation
-CO @ 0.615 µmol h−1
CH4 @ 0.063 µmol h−1
0.140 b
CdS–WO3
heterostructure
[86]
simple precipitation methodblown with nitrogen for 30 min300 W Xe arc lamp with a UV-cutoff filter (λ ≥ 420 nm);
6.0 mW cm−2
H2O100 mg catalyst +10 mL of distilled water to form films; 200 mL reactor; 1 mL sample gas extracted every 1 h;
AQY @ λ = 420 nm; atmospheric pressure and ambient temperature
-CH4 @ 1.02
µmol g−1 h−1
0.40 b
CeOx-S/ZnIn2S4
hybrid structure
[87]
one-pot hydrothermal methodintroduction of high purity CO2 gas into the reactor for 3 min9.0 W
(455 nm LEDs)
0.5 mL H2O10 mg catalyst; 6.98 mL reactor; 1 bar; below 42 °C; 10 h irradiation-CO @ 0.18
mmol g−1 h−1
1.34 b
Pt/TiO2
[42]
vacuum impregnationreactor was cleaned with nitrogen for half an hour then it was replaced and saturated with CO2 gas for at least 30 min300 W UV light;
10 mW cm−2
2 mL H2O100 mL reactor; sample gas analyzed every 1 h; 0.1 MPa; 7 h irradiation-CH4 @ 20.55
µmol g−1
10.03 b
a = AQY computed; b = AQY was computed by authors of the reference paper; c = AQY was computed by authors of the reference paper multiplying it with mass of photocatalyst.
Table 3. Summary of operating parameters used by other research works for photocatalytic CO2 reduction using Continuous flow reactor.
Table 3. Summary of operating parameters used by other research works for photocatalytic CO2 reduction using Continuous flow reactor.
PhotocatalystSynthesis MethodPre-Treatment of ReactorLight SourceReducing AgentReaction ParametersReactorMain ProductAQY (%)
TiO2/NRGO-300
nanocomposites
[88]
one-step urea-assisted hydrothermal methodpurged with CO2 at 16 mL min−1
for 40 min
400 W Xe
lamp (250 < λ < 400 nm);
11.5 mW cm−2
H2O10 mg catalyst; CO2 flowrate @ 3 mL min−1;
sample gas extracted every 1 h; 8 h irradiation
Continuous flow reactorCO @ 356.5 µmol g−10.0072 b,c
5GO–OTiO2
(UV light)
hybrid heterostructure
[46]
facile wet chemical impregnation techniquepurged with wet CO2 at 30 mL min−1 for 30 min500 W Xe arc lamp with a UV filter (λ > 400 nm);
81.0 mW cm−2
H2OCO2 flowrate @ 5 mL min−1; Quartz column reactor (ID = 9 mm, OD = 11 mm, length = 250 mm); sample gas extracted every 0.5 h; 1 bar; 25 ± 5 °C; 8 h irradiationContinuous flow reactorCH4 @ 2.7
µmol g−1
0.0103 b
TiO2
nanofibers
[89]
sol-gel method and electrospinning techniquefirstly, degassed under vacuum and then purged with Ar for 1 h, then fed with CO2/H2O mixture in dark for 1 h, then reactor was pressurized and kept at a reaction flow rate of 2 mL min−1 for another 1 h.four 6 W UV lamps
max = 365 nm)
H2O100 mg catalyst; 190 mL reactor; 7.25 CO2:H2O molar ratio; sample gas analyzed every 22 min; 2 bars; 50 °C; 20 h irradiationContinuous flow reactorCO @ 203.91
μmol gcat−1
0.04 b
Cu/GO-2
hybrid structure
[90]
one-pot microwave processpurged with nitrogen gas for 1 h then followed by pure CO2 for another 1 h300 W halogen lamp;
100 mW cm−2
H2O100 mg catalyst; 300 mL reactor; CO2 flowrate @
4 µL/min; 25.0 ± 0.5 °C; 2 h irradiation
Continuous flow reactorCH3OH @ 2.94µmol g−1 h−1 (AQYCH3OH = 0.0296)
CH3CHO @ 3.88
µmol g−1 h−1
(AQYCH3CHO = 0.065)
0.095 a
G/TiO2-001/101
nanocomposites
[91]
one-pot solvothermal methodpurged with the CO2 + H2O mixture at 200 mL min−1 for 1 h and then reduced to 5 mL min−1 for 30 min300W Xe arc lamp (300 < λ < 400 nm);
20.5 mW cm−2
5 mL DI water10 mg catalyst;
85 mL reactor; sample gas analyzed every 30 min; atmospheric pressure; 120 °C; 4 h irradiation
Continuous flow reactorCO @ 70.8
µmol g−1 h−1
0.0864 b,c
BWO-OV/BiOI
binanosheets
[92]
simple self-assembly approachpurged with the CO2/H2O gas mixture at 50 mL min−1 for 30 min500 W Xenon arc lamp with UV cut-off filter
(to remove λ < 400 nm)
H2OCO2 flowrate @ 5 mL min−1; sample gas analyzed every 1 h; atmospheric pressure and ambient temperature;
8 h irradiation
Continuous flow reactorCO @ 320.19 µmol g−1
CH4 @ 18.32 µmol g−1
0.432 b
Pt2+–Pt0/TiO2
nanoparticles
[93]
sol–gel methodpurged with CO2 + H2O mixture at 200 mL min−1 for 1 h and then at 3 mL min−1 for another 30 min.300 W Xe arc lamp UV light irradiation
(320 < λ < 420 nm);
32.5 mW cm−2
H2O200 mg catalyst; 85 mL reactor; sample gas extracted every 40 min; 50 °C;
7 h irradiation
Continuous flow reactorCH4 @ 264.5 µmol g−1
(AQYCH4 = 1.35)
CO
(AQYCO = 0.07)
1.42 b
(Pt/TiO2) @rGO
core-shell-structured
[94]
hydrothermal methodvacuum-treated, then purged with CO2 gas @ 50 cm3 min−1 for 30 min300 W Xe lamp (320 < λ < 780 nm);
80 mW cm−2
2.0 mL H2O100 mg catalyst; sample gas extracted every 1 h; 0.1 MPa; 4 °C; 8 h irradiationContinuous flow reactorCH4 @ 41.3 µmol g−1 h−11.93 b,c
NiO/Ni-GR
nanoparticles
[95]
pyrolysis and incipient wetness impregnationphotoreactor was heated at different temperatures300 W Xe lamp;
2236 W m−2
H240 mg catalyst; 51 mL reactor; 1.3 bar; 200 °C;
2 h irradiation
Continuous flow reactorCH4 @ 642 µmol gNi−1 h−11.98 b
Pt-TiO2
nanostructured films
[96]
aerosol chemical vapor depositionpurged with CO2 and water vapor at 100 mL min−1 for 1 h, and then reduced and maintained at 3 mL min−1400 W Xe lamp
(250 < λ < 388 nm);
19.6 mW cm−2
H2O0.7 mg catalyst; atmospheric pressure and room temperature;
5 h irradiation
Continuous flow reactorCH4 @ 1361 µmol g−1 h−1 (AQYCH4 = 2.33)
CO @ 179.34 µmol g−1 h−1
(AQYCO = 0.077)
2.41 b,c
a = AQY computed; b = AQY was computed by authors of the reference paper; c = AQY was computed by authors of the reference paper multiplying it with mass of photocatalyst.

Share and Cite

MDPI and ACS Style

Ali, S.; Flores, M.C.; Razzaq, A.; Sorcar, S.; Hiragond, C.B.; Kim, H.R.; Park, Y.H.; Hwang, Y.; Kim, H.S.; Kim, H.; et al. Gas Phase Photocatalytic CO2 Reduction, “A Brief Overview for Benchmarking”. Catalysts 2019, 9, 727. https://doi.org/10.3390/catal9090727

AMA Style

Ali S, Flores MC, Razzaq A, Sorcar S, Hiragond CB, Kim HR, Park YH, Hwang Y, Kim HS, Kim H, et al. Gas Phase Photocatalytic CO2 Reduction, “A Brief Overview for Benchmarking”. Catalysts. 2019; 9(9):727. https://doi.org/10.3390/catal9090727

Chicago/Turabian Style

Ali, Shahzad, Monica Claire Flores, Abdul Razzaq, Saurav Sorcar, Chaitanya B. Hiragond, Hye Rim Kim, Young Ho Park, Yunju Hwang, Hong Soo Kim, Hwapyong Kim, and et al. 2019. "Gas Phase Photocatalytic CO2 Reduction, “A Brief Overview for Benchmarking”" Catalysts 9, no. 9: 727. https://doi.org/10.3390/catal9090727

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