Next Article in Journal
Signaling the Adoption of the Benefit Corporation Model: A Step towards Transparency
Next Article in Special Issue
Monitoring and Assessment of Groundwater Quality at Landfill Sites: Selected Case Studies of Poland and the Czech Republic
Previous Article in Journal
Parameter Extraction of Three Diode Solar Photovoltaic Model Using Improved Grey Wolf Optimizer
Previous Article in Special Issue
Effect of Acetic Acid on Compressive Strength and Geometric Texture of the Surface of C20/25 Class Concrete
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Carbon Dioxide Conversion: A Circular Bioeconomy Perspective

1
Department of Chemical Engineering, Yuan Ze University, Taoyuan City 32003, Taiwan
2
Department of Marine Environmental Engineering, National Kaohsiung University of Science and Technology, Kaohsiung City 811213, Taiwan
3
Department of Chemical Engineering, National Cheng Kung University, Tainan City 701, Taiwan
4
Department of Civil and Environmental Engineering, University of Delaware, Newark, DE 19716, USA
5
Department of Chemical and Materials Engineering, College of Engineering, Tunghai University, Taichung City 407224, Taiwan
6
School of Environmental Engineering, University of Seoul, Seoul 02504, Korea
*
Author to whom correspondence should be addressed.
Sustainability 2021, 13(12), 6962; https://0-doi-org.brum.beds.ac.uk/10.3390/su13126962
Submission received: 9 April 2021 / Revised: 16 June 2021 / Accepted: 16 June 2021 / Published: 21 June 2021
(This article belongs to the Special Issue Strategies toward a Green Deal and Circular Economy)

Abstract

:
Managing the concentration of atmospheric CO2 requires a multifaceted engineering strategy, which remains a highly challenging task. Reducing atmospheric CO2 (CO2R) by converting it to value-added chemicals in a carbon neutral footprint manner must be the ultimate goal. The latest progress in CO2R through either abiotic (artificial catalysts) or biotic (natural enzymes) processes is reviewed herein. Abiotic CO2R can be conducted in the aqueous phase that usually leads to the formation of a mixture of CO, formic acid, and hydrogen. By contrast, a wide spectrum of hydrocarbon species is often observed by abiotic CO2R in the gaseous phase. On the other hand, biotic CO2R is often conducted in the aqueous phase and a wide spectrum of value-added chemicals are obtained. Key to the success of the abiotic process is understanding the surface chemistry of catalysts, which significantly governs the reactivity and selectivity of CO2R. However, in biotic CO2R, operation conditions and reactor design are crucial to reaching a neutral carbon footprint. Future research needs to look toward neutral or even negative carbon footprint CO2R processes. Having a deep insight into the scientific and technological aspect of both abiotic and biotic CO2R would advance in designing efficient catalysts and microalgae farming systems. Integrating the abiotic and biotic CO2R such as microbial fuel cells further diversifies the spectrum of CO2R.

1. Introduction

Controlling atmospheric CO2 concentration is essential to the mitigation of global warming. Amongst all strategies, CO2 conversion to value-added chemicals should be a top choice. In principle, CO2 conversion can readily take place naturally (biotic) by cultivating plants or algae to absorb CO2 and artificially (abiotic) by using synthesized catalysts in a controlled system to accelerate electrochemical CO2 conversion. Photosynthesis is the most known naturally occurring CO2 conversion reaction [1]:
106   CO 2 + 16   NO 3 + HPO 4 2 + 122   H 2 O   + 18   H + { C 106 H 263 O 110 N 16 P } + 138   O 2
The forward and backward reaction in Equation (1) indicates the photosynthesis and the respiration, respectively. The significance of Equation (1), in addition to the photosynthesis and natural respiration, is the involved chemical cycles of carbon, nitrogen, and phosphorus. This further implies that the reclamation of trace nutrients is an additional benefit to biotic CO 2 conversion. By contrast, artificial CO2 conversion requires energy input since CO2 has no heat value. Recent progress in catalysis greatly reduces the energy requirement in artificial CO2 reduction (CO2R).
There are excellent reviews on the mechanistic aspects and effectiveness of CO2R catalysts [2,3,4,5,6]. Artificial CO2R can be achieved by processes, such as electrochemical [7,8,9], photochemical [10,11,12], and photoelectrochemical (PEC), by which electrons are transferred to CO2 molecules to increase their energy content [13,14]. Electrochemical CO2R is an electron transfer reaction between the cathode and the adsorbed CO2. Photochemical CO2R relies on the transfer of photogenerated electrons in the LUMO (the lowest unoccupied molecular orbital) of the catalyst to the HUMO (the highest occupied molecular orbital) of adsorbed CO2. In photoelectrochemical (PEC) CO2R, the photogenerated electrons are transported to the cathode in the presence of a bias electric field. The external electric field effectively suppresses the recombination loss of photogenerated charge carriers and enhances CO2R efficiency.
In all modes of CO2R systems, catalysts (in the form of photosensitive or electric conductive) are needed to facilitate the CO2 reduction reactions. Therefore, the stability and surface chemistry of catalysts (electrodes) become important operation parameters. For instance, it is suggested that a thin γ Al 2 O 3 overlayer could effectively stabilize the Faradaic efficiency and the partial current density of the SnO2 catalyst in an electrochemical CO2R reaction [8]. Chang et al. (2016) have reported that photogenerated holes, not electrons, are the primary cause of instability of Cu2O catalysts, which require Cu2O to be operated as dark cathodes [13]. High-CO-affinity electrocatalysts (i.e., Cr, Mn, and Fe-N-C) exhibited high carbon monoxide (CO) Faradaic efficiency. The pyridinic and hydrogenated (pyrrolic) nitrogen moieties of the carbonaceous support are active sites for CO2 adsorption [9]. Accordingly, a relatively basic surface (such as the presence of γ Al 2 O 3 overlayer and pyridinic modification) would have a positive effect on CO2R efficiency enhancement, likely through increasing the affinity of the catalyst surface toward CO2 adsorption. Defects in a catalyst would introduce coordinately unsaturated sites (i.e., active sites for molecular chemisorption) and provide spatially supply channels for energy and electron transfers in photochemical CO2R [10]. Note that among all available CO2R processes, electrochemical CO2R is probably the simplest; therefore, it is the easiest and most sustainable (relative to chemical reduction) for operations and scaling-up.
CO2R involves the consecutive transfer of one electron or hydrogen from the catalyst to CO2 and intermediates, which gradually reduces the carbon oxidation number stepwise. Only intermediates that possess even oxidation numbers are thermodynamically stable. This is why formic acid (FA, HCOOH), carbon monoxide (CO), formaldehyde (HCOH), methanol (CH3OH), and methane (CH4) are always found in CO2R reactions. This review focuses on recent advances of CO2R systems, both biotic and abiotic processes, and as far as the broad view of CO2 conversion is concerned, three significant implications are noted [15].
Mitigating greenhouse gas effect: Electrocatalytic CO2R is usually conducted with high-purity CO2 based on thermodynamics considerations. This means that additional electricity is required for concentrating CO2 feedstock; however, the majority of electricity in modern society is produced through fossil fuels. Adopting electrocatalytic CO2R for the mitigation of the greenhouse effect would instead increase atmospheric CO2 concentration [2,16]. Therefore, increasing the deployment of renewable energy would be equally important in addition to the efficiency improvement of electrocatalytic CO2R from the prospect of greenhouse gas mitigation.
Electrochemical conversion of CO2 to fuels: In order to achieve a carbon neutral footprint through electrocatalytic CO2R, produced hydrocarbons should not be fed to the combustion engines directly. Instead, they should be feed for fuel cells [15]. Again, increasing the efficiency of fuel cells for feeding CO2-derived methanol [17,18] or formic acid [19,20] is another issue of concern. Selectivity of CO2R catalysts is crucial too as it minimizes the carbon footprint required for subsequent separation and concentration of hydrocarbons produced from electrocatalytic CO2R [21].
Electrochemical conversion of CO2 to a building block: Electrocatalytic CO2R to carbonaceous fuels usually leads to a positive carbon footprint as mentioned above. However, a carbon neutral footprint could be possibly reached by converting CO2 to a building block as an alternative. This is because CO2 is a C1 building block that requires no extra energy for its production [22]. For example, carbon monoxide produced by electrocatalytic CO2R could be directly used as the source for producing phosgene [23,24]. Similarly, as-obtained methanol and formic acid could be feedstocks for reversible chemical hydrogen storage and other applications [5,25].
Apparently, converting CO2 to a building block for chemicals production could entice the pursuant of electrocatalytic CO2R. Indeed, all conceptual designs of an economically affordable CO2R business for achieving a negative carbon footprint have assumed that the CO2R catalysts could exhibit excellent reactivity and selectivity. For instance, Gai et al. (2016) performed a conceptual design of methanol production from CO2 at an industrial scale using ASPEN Plus® (Bedford, MA, USA) in which CO and H2 are produced from the electrolysis of CO2 and H2O [26]. They found that when CO2 conversion is less than 42%, the optimal methanol synthesis route is CO hydrogenation. Importantly, the achievement of a near zero carbon emission power plant is strongly built on the assumption that CO is the only intermediate and no additional energy is required for CO isolation [26]. Sun et al. (2019) have developed a 20 MWth solar–wind biodistributed energy system for simultaneously biomass cascade utilization, water resource conservation, waste heat recovery, and CO2 mitigation for hydrogen, formic acid, and grapheme production [27]. Again, in their framework, the energy efficiency is vulnerable to the compromised selectivity of electrocatalytic CO2R. Based on the above considerations, it is clear that the success of a CO2R industry strongly relies on multidiscipline cooperation and that technology is part of this. Additional bonuses such as creating jobs, building blocks for chemicals, and carbon right trading would make CO2R more sustainable. Integrating the knowledge of biotic and abiotic CO2R is another useful approach. A good example is the microbial electrosynthesis system (MES), in which the microbial is responsible for biotic CO2R, while engineering the electrode (abiotic CO2R) further improves the overall CO2R efficiency. Accordingly, recent advances in abiotic CO2R and biotic CO2R will be reviewed herein. Having a deep insight into the scientific and technological aspects of both abiotic and biotic CO2R would advance the design of efficient catalysts and the microalgae farming system. We first focus on the technology aspect of abiotic CO2R by discussing the reactivity and selectivity of CO2R catalysts and reaction mechanisms in both water and gas phases. The effect of the surface chemistry of synthesized catalysts on the reactivity and selectivity of CO2R will be addressed. This will be particularly beneficial to the rational design of high-efficient catalysts for CO2R conversion. CO2R through microalgae abstraction of CO2 is highly influenced by the bioactivity of selected microalgae. Separation and purification of various value-added chemicals obtained is another issue of concern, and further refinery of algal biomass is also included in this review. Additional considerations such as the involvement of other stakeholders that allow CO2R to be more sustainable are also discussed.

2. CO2 Conversion Processes

2.1. Homogeneous Catalysts for CO2R

Figure 1 illustrates the different reaction steps involved in the homogeneous and heterogeneous CO2R. In the homogeneous CO2R system, two reaction steps are suggested: (i) the redox of the catalyst and (ii) the charge transfer from the catalyst to CO2. The homogeneous CO2R catalyst is also known as the molecular catalyst and is designed by mimicking the structure of chlorophyll. In the chlorophyll structure, both the metal center (Lewis acid) and the chelating ligand (Lewis base) are known to have a strong influence on the energy gap and electron transfer pathway [12]. For example, the porphyrin (in porphyrin-Co, porphyrin-Cu, and porphyrin-Au homogeneous catalysts) has been identified as both a ligand and a photoswitch to regulate the electron transfer pathway to the metal center [12]. An important feature of the homogenous catalyst for CO2R is that CO2 reduction occurs at the potential of the catalyst, not at the thermodynamic potential, where the CO2 reduction takes place [2]. Accordingly, the mass transport of homogeneous catalysts from the diffusion layer to the electrode surface therefore regulates the CO2R efficiency. It is difficult to conclude whether either homogeneous or heterogeneous CO2R reduction is advantageous over the other. The CO2R efficiency in the continuous operation unit of the flow cell will be high over homogeneous catalysts. However, surface engineering at a heterogeneous catalyst for enhancing CO2R selectivity is relatively easy. The sluggish oxygen evolution kinetics occurring at an anode (as it involves a four-electron transfer) limits the overall CO2R efficiency [28].

2.2. Heterogeneous Catalysts for CO2R

In the heterogeneous CO2R system, three reaction steps are involved: (i) CO2 adsorption at the catalyst surface; (ii) charge transfer from the catalyst to adsorbed CO2; and (iii) desorption of reduced products. The important CO2R intermediate in this system is the highly reactive CO 2 radical anion, which is produced through reorganizing linear CO2 to the bent radical anion [29]. Apparently, bending linear CO2 requires significant amounts of additional energy to form/break chemical bonds (chemisorption) and transfer electrons. Accordingly, kinetics in bond-forming interactions and electron transfer significantly regulate CO2R efficiency. The distribution of CO2R products is further identified to be closely influenced by the activation energy of chemisorption [2,3,4]. For example, the activation energy of metallic catalysts, such as Pt, Ag, and Cu for CO2 chemisorption is relatively low. This accounts for the fact that CO is the major product in this case as CO2 reduction is initialized by the charge transfer from the catalyst surface to the adsorbed CO2 [30]. By contrast, hydrogen transfer would be the dominant reaction in CO2R if the activation energy for CO2 chemisorption is relatively high. This explains that the CO2R through Cu cathode mainly produces hydrocarbons, methanol, and formaldehyde [31]. Nanostructured catalysts are frequently used in heterogeneous CO2R because of the large specific surface area, which always comes with a great fraction of highly active low-coordination sites, such as edges, steps, and defects [32]. Composite CO2R catalysts, such as metal–metal oxide, metal–carbon, and others will exhibit synergistic effects as they accelerate reaction kinetics, enhance catalyst stability, and improve selectivity relative to their individual counterpart [3].

3. The Chemistry of Abiotic CO2R

3.1. Abiotic CO2R in Water Phase

CO2R can occur in the water or gas phase. In the former system, carbonate species, namely, H2CO3*, HCO3, CO32−, are reduced. In the latter, gaseous CO2 is reacted with electron donors over catalysts and C1 or C2 compounds such as CO, formate, methanol, and oxalate are major products. Undoubtedly, the usage of rare and precious metals, such as Re and Pd, always leads to the highest CO2R efficiency [4]. The application of transition metals such as Fe, Mn, and Ni has received much attention recently because of material abundance and economical affordability [33,34].

3.1.1. Effect of Cu Surface Chemistry on Abiotic CO2R

Buckley et al. (2019) have studied the structure-reactivity relationships of electrocatalytic CO2R on modified Cu cathode surfaces [35]. The Cu cathode is first modified with long chain hydrocarbons so as to render the Cu surface hydrophobic. Modified electrodes are used to study the CO2R reaction in CO2-saturated KHCO3 (0.05 M) solution [36]. The Faradaic efficiency (ηF) of each species is calculated by Equation (2):
η F = nFVC Q
where F is the faraday constant (96,485 C-mol−1), V is the volume of electrolyte (specifically the catholyte), C is the concentration of carbonate species (M), and Q is the total charge passing through the cathode during electrocatalysis CO2R. n is the number of electron transfer in the CO2R process. For example, n = 2 for CO2 reduction to CO or HCOOH (FA). The partial current density (ji) of certain species is the product of the Faradaic efficiency and total current density (jtotal):
ji = ηF × jtotal
Figure 2 shows the replotted contour image of the current density of FA, CO, and H2 production in electrocatalytic CO2R over the modified Cu cathode [35]. Two interesting features are noted in Figure 2. First, FA is the major CO2R species on Cu-based cathodes as its value is about tenfold higher than that of CO. Second, the hot zone (the red area) of the hydrogen current occurs at the left-handed side of the figure that corresponds to the region with the lowest CO current. This clearly indicates that electrocatalytic hydrogen production profoundly competes with CO evolution. Importantly, the hot zone of hydrogen evolution appears in the region with a moderate FA current (0.5~2.5 mA/cm2). This means that FA formation is less affected by hydrogen evolution than that of the CO formation [13]. It must be mentioned that increasing operation potential inevitably leads to high H2 yield (Figure 3) [37]. Due to thermodynamics restrictions, FA and CO evolution in CO2R is suggested to be carried out at a relatively low potential condition.
Engineering electrochemical properties of cathode materials enhances the selectivity. Designing an efficient catalyst for CO2R is a highly technical challenge due to the strong completion from the hydrogen evolution reaction (HER) [3]. Factors such as catalytic reactivity, product selectivity, Faradaic efficiency, catalytic stability, and reduction mechanisms are crucial to controlling the efficiency of CO2R [38]. A systematical evaluation of the effect of catalyst structure on reaction selectivity is hence highly desirable [4]. Taking the electrocatalytic CO2R molecular system as an example, the energy required to dissociate an M–H bond to form a hydride is the key parameter in CO2R selectivity [2,4,39,40]. Figure 4 shows pathways regulating the transfer of either two protons (for CO evolution) or two electrons (for formic acid formation) and in both cases hydrogen evolution is always the major competitive side reaction.
In this framework, the preferential interaction between the catalytic metal center and CO2 over protons is responsible for the selectivity for CO evolution. While the moderate hydricity facilitates CO2 insertion into M–H bonds for FA production, strong hydride donors catalyze H2 formation [2,4,39,40]. It has been suggested that catalytic activity requires the presence of a weakly coordinating solvent molecule that can readily become dissociated during the catalytic cycle as to provide a vacant coordination site for water binding and assisting C−O bond cleavage [4]. Generally, H2 and FA formations are favorable reactions in aqueous solutions [39]. In fact, by plotting the hydricity as a function of individual free energy for the one-electron reduction of the parent species, a linear correlation appears indicating high FA selectivity over CO formation. Importantly, this correlation extends over a wide range of metals, ligand architectures, structural geometries, and overall charge of the metal hydride [40]. High overpotential is always found in CO2R [41]. Despite hydrogen evolution always being a competing reaction in CO2R, it is still worthy of scientific investigation on the hydrogenation of CO2 to FA and dehydrogenation of FA as a practical hydrogen storage pathway [5]. FA formation essentially increases the density of hydrogen gas [6]. This opens a practical alternative, such as the direct formic acid fuel cells (DFAFC) [42]. Major heterogeneous metal catalysts, such as In, Sn, Hg, and Pb exhibit high FA selectivity [5,6,41,42]. Electrochemical CO2R on polycrystalline Sn surfaces exhibits high FA selectivity too. Formation of *OCHO at Sn surfaces is the key intermediate for FA production due to optimal *OCHO binding energy. The results suggest that oxygen-bound intermediates are critical to understanding the mechanism of CO2 reduction to HCOO– on metal surfaces [43].
Cu is known to have relatively low CO selectivity because the CO produced is further reduced to several multi-carbon oxygenates (i.e., ethanol, acetate, and n-propanol) [41]. Specifically, sulfur-modified copper catalysts (Cu–S) exhibit positive correlation between particle size and selectivity toward FA evolution [44]. Nanostructured porous dendritic Cu-based catalysts show stable and selective conversion of CO2 into FA at high current density with low overpotential [45]. The relatively low CO selectivity on Cu surfaces results from consecutive CO electroreduction activity [46]. That is, in CO2-saturated aqueous solutions, polycrystalline Cu catalysts produce a mixture of compounds. Indeed, H2 evolution is dominated at low overpotential, CO and FA formation mainly occurs at high overpotential, while hydrocarbons, ethanol, acetate, and n-propanol formation happen at the most extreme overpotentials [47,48]. In a CO2 free environment, CO is reduced to hydrocarbons and multi-carbon oxygenates over the Cu catalyst [49,50]. Interestingly, oxide-derived Cu (Cu catalysts prepared by reducing Cu2O) shows much higher H2 selectivity than polycrystalline Cu [51]. Similarly, aqueous electrochemical CO reduction to C2 products by face-to-face coordinated thiol-terminated metalloporphyrins on copper electrodes exhibits 83% Faradaic efficiency and 1.34 mA/cm2 current density at −0.40 V vs. RHE. This is a significant improvement in both selectivity and activity by one order of magnitude over parent copper surfaces or copper functionalized with porphyrins in an edge-on orientation [52]. In a similar system, oxide-derived copper (OD-Cu) electrodes exhibit a high CO reduction performance by producing ethanol and acetate with >50 % Faradaic efficiency at −0.3 V vs. RHE [53].
In a short summary, the selectivity of abiotic CO2R in the aqueous phase is highly sensitive to the surface chemistry of Cu-family catalysts. CO2R occurring at the aprotic surface tends to yield CO as the major product. By contrast, CO2R happening at the protic surface is prone to produce hydrocarbons and multi-carbon oxygenates as the major products. In this case, the selectivity is strongly affected by the involved reaction pathway. In the former case, the reduction is achieved through the charge transfer, while in the latter case the hydride transfer is mainly responsible for the CO2R. The adsorption affinity between reduced intermediates and Cu-family catalysts is another critical factor regulating the CO2R selectivity. High adsorption affinity slows the desorption of the reduced intermediates, which enables their consecutive reduction. This explains that the hydrocarbons and multi-carbon oxygenates as the major products are found in this case. It is thus concluded that increasing CO2R selectivity could be achieved through the modification of surface hydrophobicity and adsorption affinity.

3.1.2. Effect of Surface Chemistry of Non Cu-Family Catalysts on Abiotic CO2R

In the iron-based CO2R, introducing the extra elemental Fe plate profoundly decreases the overpotential of the microbial electrosynthesis system (MES) [54]. In this MES system, all produced CO2 is reduced to formate at the cathode and vast hydrogen is produced during the digestion of waste activated sludge. This is attributed to the high selectivity toward formic acid evolution over CO and methane production to the reduction of H+ at the cathode due to the slow methanogensis in Fe-C MES [54]. In the case of photocatalytic CO2R, enhancing sunlight conversion efficiency is always accompanied with improving CO2R selectivity [55]. Similar to the strategy adopted in the dye-sensitized photoelectrochemical cells [56], increasing selectivity in FA formation is usually achieved through coordinating active metals with covalent organic frameworks. For example, the columnar orientation COFs (covalent organic frameworks) provides a high-efficient charge carrier transport through the ordered π-electronic pathway, which improves electron transfer from COF to metal moiety and thus increases the reactivity [55]. The results of density functional theory computations further reveal that COFs decorated with electron-donating substituents favor CO2 reduction by decreasing the hydricity of the Rh–H bond. This results in a lower hydride transfer barrier toward formic acid production [57] because the selectivity toward CO or HCOOH production is dependent on the coordination environment of the metal ion being capable of cleaving the C–O bond in the metal–CO2H intermediate [55]. Specifically, an electron-rich coordination environment breaks the C−O bond to form CO, whereas an electron-deficiency coordination environment tends to enhance the C−O bonding thereby enhancing FA formation [58]. That is, if the center metal in covalent-organic frameworks (COFs) is a strong π-donor, such as Co(II), it usually tends to promote CO evolution. By contrast, a weak π-donor, such as Zn(II), favors HCOOH production [55]. A catalyst exhibiting low adsorption energy for HCOO* (i.e., a large energy difference between the two adsorbed CO2 reduction intermediates, namely HCOO* and COOH* and large H* adsorption energy) would have high FA selectivity [59]. Ajmal et al. (2019) have studied the selectivity of the CO2 reduction reaction over bimetallic CuZn alloy catalysts and reported that the Faradaic efficiency and partial current density of FA on Cu0.5Zn0.5 (equal molar ratio of Cu and Zn) are enhanced by nearly 4 and 5 times, respectively, that of Cu foil [60]. The high selectivity of the CuZn bimetallic alloy catalyst is originated from the synergistic effect of Cu and Zn. In this case, the Zn (a weak π-donor) is likely to create a large energy difference for the adsorption of two CO2 reduction intermediates, namely, HCOO* and COOH*. An et al. (2019) have studied CO2 reduction over Sn/SnOx catalysts and reported a maximum FA Faradaic efficiency of 89% at −1.7 V (vs. Ag/AgCl) in a 0.1 M CO2-saturated KHCO3 solution [61]. The authors further concluded that Sn(IV) and Sn(II) species are mainly responsible for controlling the overpotential and suppressing H2 evolution toward improved FA selectivity. Chatterjee et al. (2019) have reported that nanoporous Pd-based alloys (np-PdX, X = Co, Ni, Cu, and Ag) exhibit FA selectivity following the order: np-PdAg > np-PdCu > Pd/C > np-PdNi > np-PdCo [62]. They have concluded that the composition-dependent behavior was governed by CO adsorption strength associated with the presence of transition metal alloying components near the Pd-skin surface and a composition-dependent change in the near surface H-sorption capacity. Interestingly, the free-standing np-PdCo and np-PdNi catalysts are able to sustain a high formate partial current density (>20 mA-cm−2) with high CO poisoning tolerance while exhibiting insignificant loss of the active area [62]. This further highlights the importance of durability and resistance of catalysts against CO poisoning during CO2R.
The results of computational hydrogen electrode model simulation reveal a striking similarity in CO2R electrocatalytic activity for the Cu3 vs. Cu5 and Cu4 vs. Cu6 size-selected clusters [63]. The rate-limiting potential of Cu4 and Cu6 clusters in CO2R is the proton-electron (H+ + e) transfer to CO* (species adsorbed on clusters) to form CHO*, which is also the rate-limiting step on Cu surfaces. On the other hand, with respect to Cu3 and Cu5 clusters, removing OH* from the cluster surface (OH* S + OH ) is the rate-limiting step in CO2R [63]. The above simulation unambiguously implies the role of surface defects, in addition to bulk electrocatalysts, in regulating CO2R reaction pathways. Indeed, the electrolysis of CO2 on 4-aminomethylbenzene-modified Pb electrodes exhibits a current density as high as 24.0 mA/cm2 (at −1.29 V vs. RHE) and a FA Faradaic efficiency greater than 80% [28]. Pt-based alloys having high-index facets generally show high specific catalytic activity over those having low-index facets [64]. Exposing the high-index facets of nanosized particles is promising to enhance Pt utilization and at the same time enriches crystalline defects in the CO2R catalyst [64]. Similarly, Pan et al. (2019) have reported that N,S-codoped carbon catalysts exhibit 92% CO Faradaic efficiency and CO current density of 2.63 mA/cm2 at a low overpotential of 0.49 V versus RHE [65]. Incorporating S in N-doped carbon introduces a high population of activate pyridinic N sites, which significantly decreases the free energy barrier for the formation of intermediate *COOH thereby enhancing CO adsorption toward high CO selectivity [65]. The high CO selectivity of the Pd85Cu15 catalyst is attributed to the presence of a larger number of low-coordination Cu sites than active monometallic Pd sites on the catalyst [66]. Accordingly, manipulating the size and chemical composition of bimetallic nanoparticles is critical to the selectivity of CO2R [66]. Results of density functional theory calculation indicate that high reactivity and selectivity are the outcome of defects that stabilize the *OCHO intermediate [67]. Surface modification of the Cu catalyst with protic, hydrophilic, and cationic hydrophobic species results in increasing the selectivity of H2, FA, and CO, respectively [35]. Table 1 summarizes the performance of various CO2R processes. Note that Faradaic efficiency alone is not sufficient to express the degree of selectivity because the current density of individual species is also an important characteristic of an efficient catalytic CO2R reaction. In addition, as mentioned above, a low faraday efficiency in CO/FA is usually accompanied by a high faraday efficiency in H2 evolution, which is another valuable product of CO2R in the aqueous solution.
Electrolytes also play a crucial role in determining CO2R selectivity. CO2R over the BDD electrode in the KClO4 electrolyte produces CO, whereas FA is the major product in the KCl electrolyte. This is because ClO4 promotes the adsorption of CO 2 intermediates [37]. Similarly, Eder et al. (2019) have reported that on Ru-based catalysts, hydrosilanes additives and KF stabilized formate intermediates (silylformate) forms form potassium formate with a turnover number of 110 mmol-formate/mmol-Ru [70]. Note again that the CO2R selectivity is a function of applied voltage, which gradually shifts from HCOO /HCOOH to CO/H2 with increasing overpotentials [68,69]. Based on the above consideration, a new approach for effective CO2R by an aluminum hydride-like reductant has been attempted [71]. The reductant is an organoaluminum complex containing a formal aluminum double bond (dialumene). Weetman et al. (2019) have demonstrated that dialumene improves the selective formation of formic acid equivalent via the dialuminum carbonate complex rather than the conventional aluminum–hydride-based cycle [71]. Likewise, Zhao et al. (2019) have reported that the KBH 4 reduces CO 2 to HCO 2 readily, accompanied by the release of activate intermediate species and H+. Further, CO2R is accelerated by a Cu/Ni bimetal catalyst that effectively regenerated the active boron species [72].
Briefly, the surface chemistry of non-Cu-family catalysts also strongly influences the selectivity of abiotic CO2R in the aqueous phase. Unlike the Cu-family catalysts, the unique characteristics of non-Cu-family catalysts is that they can be tailored with a specific porous framework. In this configuration, the kinetics in the confined space govern the CO2R selectivity. The localized coordination environment is another factor regulating the overall CO2R selectivity, which can be modified via doping or introducing defects. Along with the modification is the adjustment in the energy levels that further affects the lifetime (stability) of the COR2 intermediate and consequently the CO2R selectivity.

3.2. The Chemistry of Gaseous Phase Abiotic CO2R

Several valuable chemicals such as CO, methane, methanol, low olefins, and long-chain carbohydrates could be produced from the gaseous phase CO2R reaction. To achieve selective CO2R, catalysts that are effective in activating both H2 and CO2 and stabilizing surface intermediates are needed. To this end, most catalysts are comprised of metallic sites, which are active in splitting adsorbed H2 (from H2 to H*) and exhibit a high affinity toward CO2 adsorption. Additional modification with alkali or noble metals can further change the surface acidity or aid in the formation of an extra alloy phase in the catalyst [73]. The metal/support interfacial sites are highly active in CO2 hydrogenation due to electron perturbation of the metal and partial reduction of metal oxide via the H-spillover mechanism [74,75].
Wet impregnation and co-precipitation are the two most frequently used methods to synthesize heterogeneous catalysts for CO2 hydrogenation [76,77]. Nanosized catalysts can be synthesized by flame-spray pyrolysis [78]. The catalysts made by pyrolyzing metal organic precursors can achieve a complex nanostructure with high surface metal loadings [79,80] and the carbon sites can promote the adsorption and activation of CO2 [81]. Fixed bed reactor configuration is applied in the evaluation of catalytic performance, in which specific amounts of catalysts are distributed over inert particles, such as SiC or silica, to the minimize hot spot phenomenon. After pre-reduction of the catalysts in H2 flow, the stream that contains certain ratios of H2/CO2/inert gas with a specific gas hourly space velocity (GHSV) is injected into the fixed bed reactor. The CO2 conversion (XCO2) and selectivity (S) toward CO and low hydrocarbons are measured from on-line gas chromatography, while the long chain hydrocarbons are collected in a cold trap for further quantification.

3.2.1. Hydrogenation of CO2 to CO and CH4

Under ambient pressure, hydrogenation of CO2 over most metal catalysts produces either CO or CH4. The product CO can serve as a feedstock in the methanol synthesis process and the Fischer–Tropsch process for further carbohydrate synthesis [73]. Conversion of CO2 to methane could buffer the fluctuations in energy supply via the power-to-gas process that converts excess electricity to H2 as the reducing agent in CO2 methanation [78]. Equations (4) and (5) present the Sabatier reaction and reverse water–gas shift (RWGS), corresponding to the hydrogenation of CO2 to methane and CO, respectively. Thermodynamically, the former reaction dominates at T < 500 °C, while the latter at T > 500 °C. However, as presented in Table 2, the selectivity toward CO and methane is greatly altered on the heterogeneous catalysts with the combination of various metals and supports.
CO 2 + 4 H 2     CH 4 + 2 H 2 O
ΔH298K = −165 kJ/mol
CO 2 +   H 2     CO   +   H 2 O
ΔH298K = 41 kJ/mol
In the presence of noble metals, CO2 hydrogenation to CO can be realized at a low temperature. This phenomenon is attributed to the lower activation energy of the hydrogenation process on the active sites of metals [86]. For instance, Dietz et al. (2015) have simulated RWGS at the (111) plane of several metals and found that Ni, Cu, and Rh favor the dissociation of CO2 → CO + O, while Ag, Pd, and Pt prefer the hydrogenation pathway: CO2 + H → COOH [86]. When Pt is loaded on silica and titania, the support itself enhances CO2 adsorption. The energy change of CO to HCO governs the selectivity toward CO, while the competition for *H2COH between hydrogenation and C–O bond cleavage affects the preferential production of CH4 or CH3OH [87]. The performance of Pt on RWGS is enhanced by the addition of a potassium promoter, which enables the formation of Pt-O(OH)-K interfacial intermediate that promotes the adsorption of the bicarbonate species, the precursor of CO via the formate pathway. CO2 conversion in the K-promoted Pt/zeolite system at 500 °C is 27.4%, which is 2-fold greater than the system without K-promoters [88]. Kattel et al. (2016) further reported that the interfacial sites between PtCo alloy and other reducible oxides (CeO2, TiO2 and ZrO2) are important to stabilizing surface intermediates [82]. Wang et al. (2015) have studied the mechanisms of CO2 hydrogenation over Pd/Al2O3 and concluded that RWGS and the Sabatier reaction do not take place at the same surface sites [89].
Ru-based catalysts show great catalytic methanation at a low temperature. Dreyer et al. (2017) have investigated the hydrogenation of CO2 using Ru-based catalysts dispersed on different metal oxide supports, including Al2O3, ZnO, MnOx, and CeO2 [78]. They have found that methanation occurs by partial reduction at metal oxides supports, which increases the coverage of H* but strengthens the C–O bond of CO*. The highest CO2 conversion (83%) and methane selectivity (99%) at 300 °C is obtained by the Ru/CeO2 system [78]. Guo et al. (2018) have demonstrated the metal-support interactions and the effect of H-spillover on CO2 methanation [90]. By varying the degree of Ru dispersion from the size of a single atom to nanoparticle (4 nm) on Ru/CeO2, metal-support interaction is the strongest for a single-atom Ru/CeO2 that facilitates CO* activation, whereas H-spillover prevails in large Ru clusters and prevents the catalyst from poisoning by enhanced H2O removal [90]. Thus, controlling the size of Ru at around 1.2 nm achieves an appropriate balance between the two phenomena, leading to a turnover rate of 1.6-fold and 14-fold greater than single-atom and 4-nm Ru, respectively [90]. The interactions between Ru and TiO2 are fortified by syngas pretreatment at 600 °C, which leads to an increase in interfacial sites on Ru-TiO2 by encapsulation of Ru [91]. Hydrogenation of CO2 on nickel-based catalysts has been explored extensively. Dispersing Ni on SiO2 support is deterministic of selectivity, in which the 10 wt% Ni/SiO2 is effective in stabilizing the monodentate configured HCOO intermediate, which does not occur on catalysts of low Ni loadings (0.5 wt%) [92]. Bi et al. (2019) have demonstrated that impregnating Ni on an MCM zeolite with a sodium-free alkaline agent enhances the synergism between Ni and NiO during CO2 hydrogenation via H2 adsorption and CO2 activation, respectively [85]. The catalyst exhibits remarkable CO2 conversion (68.3%) and methane selectivity (91.4%) with high stability [85]. Doping Ni/ZrO2 with iron enhances the reducibility of Ni and ZrO2 owning to the electron-donating property of Fe(II), which in turn promotes the synergism effects of Ni-NiO and metal-support interactions [83,93].
Based on the above discussions, it is noted that in addition to the reactivity of CO2R catalysts, the chemical environment of the support also plays an important role in CO2R selectivity in the gaseous phase. This is because the gaseous abiotic OC2R is conducted in a high temperature (in comparison with the condition of the aqueous CO2R). In this configuration, the dispersion of CO2R catalysts and consequently their contact with support is highly sensitive to the stability of support. A relative alkaline support such as alumina is beneficial for CO2R efficiency as CO 2 is regarded as a weak acid in this configuration. It is thus suggested that in addition to the characteristics of the CO2R catalyst, its dispersion over the support is another issue of concern for rationally engineering a CO2R catalyst with high selectivity and efficiency.

3.2.2. Hydrogenation of CO2 to Methanol

Methanol is a valuable CO2 reduction product as it can be served as biofuel, building blocks in organic synthesis, and fuels for a methanol-based fuel cell. Table 3 shows CO2 conversion to methanol at a pressure of 1–4 MPa over different metallic meal oxide catalysts. Catalyst Cu/Zn/Al2O3 plays an important role in the commercial methanol production from syngas. DFT studies reveal that the stabilization of the transition surface species is the key to achieving high selectivity in the syngas conversion. The formate pathway predominates the total process on the surface of Cu/ZnAl2O4 and Cu/Zn/Al2O3, in which the rate-limiting step is the formation of H2COO* and H2COOH* [94,95]. The CO production via RWGS is the major byproduct in methanol synthesis. Based on the H/D isotope substitution technique, it is known that methanol synthesis and RWGS occurrence take place at difference surface sites on Cu/ZnO/Al2O3, Cu/MgO, Cu/SiO2, and Pd/SiO2 [96]. Karelovic and Ruiz (2015) have suggested that ZnO loads with larger Cu particles tend to suppress the activity of RWGS as the specific methanol formation rate per surface Cu is independent of Cu particle size, while that of CO is enhanced by smaller Cu particles [97]. Ro et al. (2016) have studied the synergistic effect by dispersing Cu on ZrO2 and reported that the rate constant of CO2 conversion on Cu-ZrO2 interfacial is eight times greater than that on plain Cu [98]. Phongamwong et al. (2017) have added colloidal silica on Cu/Zn/ZrO2 as geometric spacers to enhance the stability and performance of the Cu-based catalysts in CO2 hydrogenation [99]. The authors have reported that 1% of SiO2 loading increases the methanol synthesis activity by 26% and retains 12% more activity after emerging from the steam for 96 h at 280 °C.
Bimetallic Pd, such as PdZn and PdCu, exhibits a high methanol yield similar to Cu-based catalysts even at low temperatures [73]. The Pd dispersed in ZnO shows exceptional stability attributed to the particle size being maintained at 5 nm even after pre-reduction at 400 °C. While the colloidal dispersion technique is effective in stabilizing the interfacial sites of PdZn, its methanol formation yield is 40-fold greater than the same catalysts synthesized by the traditional wet impregnation method [101]. PdCu and PdCu3 bimetallic catalysts exhibit a methanol formation rate of 0.31 μmol gcat−1 s−1, which is 3.4-fold and 6.2-fold greater than monometallic Pd and Cu, respectively [100]. Results of DFT simulation further indicates that the (111) plane of PdCu is highly active in methanol evolution, particularly at the low-coordinated Pd on the stepped surface [102]. Furthermore, methanol evolution through the formate pathway again is the major catalytic reaction, which can be further promoted by adsorbing small amounts of water to lower the energy barrier through the H-shuttled mechanism [102]. A novel metal-free In2O3 catalyst is also known to exhibit high selectivity in reducing CO2 to methanol [76]. The oxygen vacancies on In2O3 are important to CO2 hydrogenation. Specifically, the In2O3 with optimal oxygen vacancy exhibits a methanol yield of 3.7 mol kg-cat−1 h−1, CO2 conversion of 7.1%, and methanol selectivity of 40%, respectively, at 330 °C and 4 MPa [75]. Martin et al. (2016) have reported that the selectivity toward methanol over In2O3 approaches 100% in the temperature range of 200–300 °C [104]. The methanol yield is further promoted by increasing the density of surface oxygen vacancy through Ar sputtering, syngas pretreatment, and ZrO2 support [104]. Likewise, incorporating a small fraction of Mn into the spinel Co3O4 structure greatly enhances the methanol selectivity in CO2 hydrogenation, likely attributed to the increase in surface basicity [103]. In brief, the significance of the thermal stability of the support in the CO2R to methanol is relatively less profound than that in the CO2R to CO and methane. This is because the former is usually carried out in a relatively lower temperature than that of the latter. In this case, the surface basicity becomes much more significant in affecting the reactivity of CO2R catalysts.

3.2.3. Hydrogenation of CO2 to Low and Long Chain Chemicals

CO2 conversion to lower olefins (C2–C4), building blocks, and other long chain hydrocarbons in the gasoline range (C5–C11) or diesel range (C12–C21) has been explored extensively. Again, catalysts that are reactive in both CO2 hydrogenation and the Fisher–Tropsch reaction will be capable of achieving the CO2 conversion objectives [105]. Table 4 lists the performance of CO2 hydrogenation to low and long chain hydrocarbons on Fe-based catalysts. Upon CO2 hydrogenation to different hydrocarbons, the Fe-based catalysts undergo consecutive phase transitions together with the creation of multivalent charges and this surface reconstruction process further diversifies distinct active sites on the catalyst [84]. The creation of multivalent charges results in the spinel iron (Fe3O4) being composed of Fe3O4, iron carbides, and α-iron, which enhance the activity of RWGS, carbon chain growth, and olefins’ secondary reactions [106]. It is therefore suggested that an appropriate fraction of Fe3O4 and Fe5C2 is necessary for the production of high olefin and paraffin [107,108]. Pretreatment of Co-Fe bimetallic under various reducing gases (i.e, H2, syngas, and CO) is another effective ex-situ modification strategy to precisely control the phase transition [77]. Under the selected reducing environment, CO activation leads to the formation of CoFe alloy and carburized phases (Co2C and FeCx) that shift the selectivity toward low hydrocarbons and oxygenate [77]. While most Fe-based catalysts produce lower hydrocarbons, delafossite (CuFeO2) exhibits high selectivity toward long chain hydrocarbons (C5+) and 85 wt% of the produced long chain hydrocarbons are in the gasoline and diesel range [109]. Doping alkali metals on Fe-based catalysts is also beneficial to chain growth propagation due to stronger COx adsorption [106]. Compared to Na-free Fe3O4, incorporating 1.18 wt% of Na greatly enhances the CO2 conversion and selectivity toward light olefin from 29.3 to 40.5% and 0.1 to 40.3%, respectively [110]. The presence of potassium promoter on Fe-Co/Al2O3 diminishes the density of hydrogen on metal surfaces, which in turn suppresses the hydrogenation of olefins [111].
Bifunctional catalysts have emerged actively in synthesizing hydrocarbons from CO2 with great flexibility. Combined In2O3 and zeolite (HZSM-5) effectively synthesizes liquid fuels (C5+) from CO2 by suppressing undesired C1 products [112]. In the process, CO2 is first hydrogenated to methanol at the reduced site on In2O3. Methanol that diffuses to the acidic site of zeolite transforms to hydrocarbons by the hydrocarbon-pool mechanism, which results in 78.6% of C5+ hydrocarbons [112]. Replacing the zeolite HZSM-5 by SAPO-34 increases the selectivity toward lower olefins, probably as a result of changes in topology [74]. Notably, the bifunctional catalyst must be packed in granular instead of powder form as the surface acidity of the zeolite support is another important factor controlling the selectivity of CO2 hydrogenation toward hydrocarbons [112]. HZSM-5/Na-Fe3O4 composite also effectively converts CO2 to hydrocarbons in the gasoline range [110]. The above catalyst composites provide three distinct active sites: Fe3O4 sites for RWGS, Fe5C2 sites for FTS, and zeolites for oligomerization [105]. Na-Fe3O4/HZSM shows little deactivation; the CO2 conversion and C5+ selectivity remain at 27% and 54%, respectively, over 1000 h of operation [105]. Based on the above discussions, it is noted that the surface chemistry of the support is of equal importance to catalysts in CO2R. Interestingly, surface acidity instead of surface basicity is much more important in this case as the development of long chain chemicals strongly relies on the hydride transfer.
In short, there are advances in the development of heterogeneous catalysts for CO2 hydrogenation in recent years due to a better understanding of the elementary reactions via DFT simulation, catalytic performance of metal-support, and metal-dopant interfacial sites. With appropriate modifications, CO2 can be hydrogenated to CO, CH4, CH3OH, and other low and long chain hydrocarbons. Further challenges are reducing cost and increasing durability of catalysts, reducing cost of renewable hydrogen sources, and carbon dioxide capture.

4. Biotic CO2R

Biological CO2R has been around for a long period of time. CO2R by microalgae cultivation takes additional advantage over terrestrial plants because the photosynthesis efficiency of the former is about 10~50 times greater than that of the latter system [113]. The additional advantage of microalgae cultivation over other higher plants lies in this capability of being spatially cultivatable. Spatial arrangement can effectively maximize the land usage in accordance with the theme of the UN Sustainable Development Goals in which the balance among the natural science, social science, and stakeholders is particularly emphasized. Research activities, such as screening for appropriate microbial species, developing a novel cultivation strategy, and separating value-add byproducts have attracted much attention. Nutrient removal and recovery from nutrient-rich wastewater is an additional bonus of microalgal CO2R [114]. Sunlight, algal, carbon source, and nutrients are four essential components in microalgae CO2R as shown in Equation (1), which can be operated in open (outdoor with ponds and raceway reactors) or closed systems (indoor with columns) [113,115]. The crucial component in biological CO2R facilities is the photobioreactors (PBRs), which can be assembled in a variety of configurations, such as tubular, flat-plate, bubble column, and airlift PBR. CO2 bubbles are injected at the bottom of PBR to facilitate gas–liquid transport [116]. Operation variables include gas aeration rate (volumetric gas flow rate per unit volume of culture medium, vvm), biomass productivity (dry cell weight, DCW, per liter of medium per day, (g-DCW L−1 d−1)), specific growth (d−1), and CO2 sequestration/fixation rate (g-CO2 L−1 d−1) [117,118].

4.1. Biotic CO2R by Microalgae Farming

Table 5 summarizes the performance of different algal species in biological CO2R. Chlorella is a genus of green algae that has been widely studied due to its fast growth rate. In the Chlorella family, Chlorella sp. and Chlorella vulgaris exhibit the highest CO2R rates and highest lipid productivity (precursor of biodiesel) over others [119,120,121]. Scenedesmus sp., a microalga isolated from a marble mining site, is active at a high CO2 stream (~15% CO2) [122]. Asterarcys quadricellulare and Chlorella sorokiniana isolated from water bodies near a steel plant, survive at a high CO2 concentration (15%) and temperature (40 °C) [123]. Yun et al. (2016) isolated Acutodesmus obliquus from an acid mine drainage site that remains active in an acidic environment [124]. Microalgae collected from a domestic wastewater treatment plant (WWTP) remains active even under 50% CO2 (without aeration) in pH 7 to pH 11 [125]. In addition to naturally occurring microalgae, genetic engineered Chlorella sp. is able to survive in an alkaline environment (pH 6~11) with a biomass productivity 12-fold higher than the wild type [126]. Introducing an aldolase gene of cyanobacteria into the chloroplast of Chlorella vulgaris increases the biological CO2R efficiency by 1.2 times that of the original strain [127]. Li et al. (2015) have reported that genetic engineered Chlorella sp. is tolerable to high CO2 concentration and exhibits a biomass yield of 4.06 and 3.68 g-L−1 in 1 and 30% CO2, respectively, which are 1.3-fold and 3-fold that of the original strain [128]. The mixed culture of nitrifying bacteria and Chlorella exhibits a synergetic effect that enables biological CO2R at relatively high dissolved oxygen while achieving high biomass production (23%) [129].
In light of the cultivation condition, it is known that factors including pH, light intensity, temperature, level and form of carbon, nitrogen, and phosphate and trace elements significantly affect the efficiency and biomass composition of biological CO2R. The microalgae cultivation is often conducted at neutral pH (ca. 7~9). Acidification occurs mainly by dissolution of NOx and SOx and to a lesser extent CO2, and complicates the biological CO2R, especially when flue gas is the sole carbon source. The addition of pH buffers, such as phosphates [130], HEPES [123] and bicarbonate [131] minimizes pH fluctuation. An intermittent gas injection strategy based on pH feedback precisely controls the pH of the growth medium within a 0.5 unit [132]. Vo et al. (2018) have studied the performance of Chlorella sp. and reported that the highest CO2 fixation rate (1.65 g-CO2 L−1 d−1) is an N:P ratio of 15:1 [133]. Domestic wastewater [134,135], aquaculture wastewater [120,136], and landfill leachate [137] can serve as nutrient sources for microalgae cultivation to lower operational costs. Liu et al. (2017) have reported that Chlorella vulgaris, cultivated using WWTP effluent discharge, consumes ammonia then nitrate [134]. On the other hand, Fu et al. (2019) have reports that feeding urea to Chlorella vulgaris instead of nitrate as a nitrogen source increases the biomass productivity by 14%, and a low dose of D-glucose could stimulate its photoautotrophic biomass production [138].
CO2 fixation efficiency and biomass productivity are highly related to CO2 concentration in gas aeration and the injection modes of bioreactors. Posadas et al. (2015) have showed that the metabolic pathway of Chlorella vulgaris is sensitive to dissolved inorganic carbon concentrations [135]. Gonçalves et al. (2016) have studied the effect of CO2 concentration on biological CO2R of algae, i.e., Chlorella vulgaris and Pseudokirchneriella subcapitata, and cyanobacteria, i.e., Synechocystissalina and Microcystis aeruginosa and reported that the optimal CO2 concentration is 5.35 ± 0.34% [118]. Intermittent aeration of 10% CO2 to mutated Chlorella sp. achieves a CO2 utilization efficiency of 20.4%, which is 5-fold greater than continuous aeration. The rate of CO2 biosquestration in the continuous mode is greater than that of the injection mode (1.61 vs. 1.33 g d−1) [126]. The rate of biomass production and CO2 biosequestration are closely proportional to light intensity. However, algal growth at different life stages requires a different spectrum and intensity of light. Dineshkumar et al. (2016) have reported that the light intensity of 50~125 μmol m−2 s−1 is enough to sustain Chlorella minutissima initially (~24 h) [139]. Xie et al. (2014) have demonstrated that increasing light intensity on Desmodesmus sp. improves CO2 biofixation rate [117].

4.2. The Unit Processes for Biotic CO2R

Common microalgae farming systems includes batch, fed-batch, and semi-continuous cultivation. Although batch cultivation is easy to operate, nutrient depletion and declining light penetrability owing to biomass accumulation gradually loses its efficiency. Fed-batch supplies nutrients periodically to maintain necessary nutrient levels. For example, Xie et al. (2014) have reported that nitrate concentration at 2 mM is necessary to maximize the productivity of lutein by Desmodesmus sp. in the batch reactor [117]. Semi-continuous cultivation, which replaces certain amount of microalgal suspension with fresh medium in order to replenish nutrients and reduce the concentration of biomass, is probably the most common strategy in large-scale algae cultivation [126,144]. Furthermore, Huang et al. (2016) have studied a pre-harvesting cultivation strategy, using filtration to decrease the concentration of microalgal cells, and reported that nitrogen utilization efficiency can reach 76%, which is 1.7-fold that of semi-continuous cultivation [143].
Biological CO2R using flue gas is a highly desirable approach as it directly reduces the CO2 emission to the atmosphere. In general, flue gas contains 5~15% CO2 in addition to certain amounts of NOx, SOx, and H2S, which is somewhat harsh to algal growth. A mixed microalgal community collected from a freshwater lake can survive in 100% flue gas (11.2% CO2, 206 ppm NO2, and 273 ppm SO2) [130]. Cultivating Scenedesmus dimorphus with synthetic flue gas containing 15% CO2, 400 ppm SO2, and 300 ppm NO under continuous aeration can suffer a sudden pH drop that eventually suppresses algal growth. Under such a circumstance, there is the intermittent sparge of flue gas damping pH fluctuation that gives rise to a high productivity of algal biomass and CO2 removal [132]. The presence of NOx in the aeration gas is beneficial for microalgae in lipid productivity [123]. Sulphate accumulation up to 870 ppm resulting from the medium recycle has limited influence on CO2R by Tetraselmi ssuecica; however, the presence of H2S inhibits the growth of Chlorella sorokiniana [121,144]. In contrast, flue gas from a nature-gas boiler containing 8% CO2 and being purged continuously at 0.05 vvm exhibits limited influence on Chlorella sp. growth [136]. Kuo et al. (2016) have reported that algal biomass production in CO2 biofixation using Chlorella sp. in large-scale bubble column PBRs is the same as that of indoor laboratory scale reactor using flue gas streams that contain 24~26%, 10~80 ppm, and 15~90 ppm of CO2, NOx, and SO2, respectively [136].

4.3. Additional Ecomomic Benefit of Biotic CO2R

The economic benefits of microalgae CO2R are unsettling. Ou-Yang et al. (2018) have argued that carbon sequestration by microalgae is not profitable for coal-fired power plants even with government subsidies and the consideration of carbon price [145]. Zhu et al. (2016) have evaluated the economy of Chlorella zofingiensis cultivation in southern China and reported that growing algae with nutrient-rich wastewater significantly reduces the cost of biomass by $451 tonne−1 from savings of water and nutrients, when the credit of wastewater treatment at $0.55 m−3 is counted [146]. The production of one kg of biomass and one L of algal oil cost approximately $1.2 kg−1 and $3.9 L−1, respectively. Large-scale microalgae cultivation for biofixation of CO2 is capital intensive and requires massive land space [147]. Therefore, the cost of the microalgae growth facility has tremendous impacts on the economy of CO2 biofixation. At a photosynthetic efficiency of 4% and target CO2 fixation of 40%, the make-even market price of microalgae is at least $440 tonne−1. Therefore, maximizing the revenue of microalgae is vital for biofixation technology from an economic perspective. There are attempts to assess the microalgae economics in terms of algae farming toward energy production, e.g., the unit area of algal biomass production and CO2 fixation. As shown in Table 5, among a total of 29 references reviewed, the biomass production rate varies by 3 orders of magnitude (0.07–2.06 g-DCW L−1 d−1), whereas some specific growth rate (d−1) data are unavailable. Without definitively and reliably consistent data, it is beyond a reasonable challenge to examine the economic feasibility of algal farming. Apart from the economic feasibility assessment, it is certain that for those existing microalgae farming pilot plants, optimization of the unit process is a top priority as it makes the microalgae farming one step closer to the market-driven business.

4.4. The Utilization of Microalgae Biomass

The conversion of microalgae biomass to valuable products, i.e., biorefinery, is key to overcoming the economical obstacle toward CO2-derived bioeconomy [148]. Figure 5 depicts the concept of biotic CO2R coupled with further refinery of microalgal biomass to a variety of products, such as biofuel, pigments, health supplements, and pharmaceuticals [149]. As mentioned above, a market-driven CO2R business will definitely be more sustainable. Based on this consideration, CO2R with flue CO2 will have a strong incentive within the industry as it eventually reduces the carbon footprint in response to the Paris Climate Agreement. Micronutrient reclamation in microalgae farming further allows the biotic CO2R to be integrated with the wastewater treatment system of the agriculture sector [150]. The interest in microalgae biomass as a renewable and sustainable feedstock for biofuels (biohydrogen, biomethane, biodiesel) and various bioproducts has inspired a new focus in biorefinery [151]. Biofuel, pigments, health supplements, and pharmaceuticals derived from the biomass harvest not only give additional economic incentives but also allows for value-added green chemicals.
Khan et al. (2018) have reported that employing microalgae to convert atmospheric CO2 to useful products such as carbohydrates, lipids, and other valuable bioproducts through photosynthesis, essentially achieves environmental and economic sustainability [152]. Algae also need N and P as nutrients, which account for 10–20% of algae biomass [153]. These nutrients can be applied as biofertilizer and redeployed in agricultural lands, and the reduction of greenhouse gas (GHG) emissions, which enables meeting the current challenges toward a circular bioeconomy. To meet the increasing food demands, microalgae are potential alternatives to animal and aquatic feeds (i.e., millet, grams, and small fishes) because of the possibility for rapid and eco-friendly cultivation, higher protein content, omega 3 fatty acids, and carotenoids [154,155].
In addition to direct food consumption, microalgae can be feedstock for further conversion to various chemicals and fuels through thermochemical processes, such as liquefication, gasification, and pyrolysis. Liquefication of microalgae takes place in sub/supercritical water, which omits the energy-intensive drying process. The yield and selectivity during thermochemical liquefication depend on the chemical composition of biomass. The application of sodium carbonate as the homogeneous catalyst promotes a bio-oil yield of carbohydrate-rich microalgae at higher temperatures (300~350 °C) and a bio-oil yield of protein-rich microalgae at low temperatures (250 °C) [156]. Gai et al. (2015) have detailed the reaction pathways of hydrocarbons, proteins, and lipids during hydrothermal liquefaction of Chlorella pyrenoidosa and Spirulina platensis [157]. Liquefying microalgae over Co/CNTs catalysts improves the biomass conversion and bio-oil yield and enhances the selectivity toward hydrocarbons [158]. On the other hand, the liquid phase produced in hydrothermal liquefication is rich in phenolic compounds with high nitrogen content [157], which can be further applied to cultivate Chlorella vulgaris and Nannochloropsis gaditana [159].
Gasification converts carbonaceous substances in microalgae to combustible gases, such as syngas and methane. Conventional gasification, which partially oxidizes dry microalgae under 800~1000°C and 1~10 bar, is a mature technology but suffers from low thermal efficiency [160]. Raheem et al. (2017) have explored the feasibility of microalgae gasification using air as a gasifying agent and reported that hydrogen gas content is decreased from 24.2 to 19.5% when the loading of biomass is increased from 1 to 2 g in the quartz reactor at the optimal temperature of 95 °C [161]. Supercritical water gasification (SCWG) enables oxidization of wet microalgal biomass in a short reaction time, in which the proper catalyst is crucial for high gas conversion and selectivity toward combustible gases [162]. A continuous SCWG of Chlorella vulgaris has been demonstrated in Switzerland to produce methane-rich gas (55~60%) [163].
Chemical looping gasification (CLG) of microalgae is a novel technology that oxidizes biomass to syngas using an oxygen carrier (MexOy, preferably iron ores) in the fuel reactor, while the reduced oxygen carrier (MexOy−1) is regenerated in the air reactor [164]. The performance of CLG on Chlorella vulgaris is enhanced by microwave pretreatment of biomass and by the adding of CaO to absorb produced CO2 in the fuel reactor [165,166]. Pyrolysis of microalgal biomass in an oxygen-free environment produces syngas, bio-oil, and char. According to the heating rate and vapor residence time, pyrolysis can be categorized into slow mode (5~10°C min−1, 10~30 s) and fast mode (10~600°C/min−1, 1~3 s) [160,166]. Wang et al. (2017) have proposed the pyrolysis pathways of carbohydrate, lipid, and protein extraction from Nannochloropsis sp [167]. To improve the biofuel and vapor quality produced from pyrolysis, catalytic pyrolysis using zeolite has received much attention. Catalytic cracking of the pyrolysis vapor over zeolite HZSM5 promotes the selectivity toward monocyclic aromatic hydrocarbons, which are important chemical building blocks [168]. Catalytic pyrolysis of Chlorella vulgaris on Ni-supported zeolite enhances the quality of bio-oil, which produces less oxygenated and acid compounds compared to non-catalytic pyrolysis [169]. Based on the above discussions, an effective catalyst straightforwardly governs the gasification efficiency. This, once again, signifies the blurry line between the abiotic and biotic CO2R. Knowledge on the rational design of efficient abiotic CO2R catalysts is essential to engineering the surface chemistry of gasification catalysts.
The biological conversion of the harvested microalgal biomass via microorganism includes anaerobic digestion (AD) and fermentation. Anaerobic digestion of the biomass produces biogas containing carbon dioxide and methane through four stages: hydrolysis, acidogenesis, acetogenesis, and ethanogenesis [170]. Fermentation converts microalgal biomass to bio-ethanol in the presence of yeast or bacteria, and to biobutanol through the classic acetone-butanol-ethanol fermentation (ABE) process [171,172]. For both cases, it is necessary to rupture microalgal cells and pretreat biomass to improve the bioavailability, anaerobic biodegradability, and fermentability. The common pretreatments are mechanical, (thermo)chemical, and biological methods. Passos et al. (2015) have reported that thermal, hydrothermal, and microwave pretreatment of microalgal biomass enhances the bio-methane yield in AD by 72, 28, and 21%, respectively, whereas those by ultrasonication exhibit no significant increase [173]. Alkali thermochemical pretreatment on microalgae with polysaccharide-based cells, e.g., Chlorella and Nannochloropsis, effectively improves the biomass biodegradability and methane production in AD mode [174]. The enzymatic pretreatment is another effective route to degrade the cell wall and solubilize cellular biomass. Passos et al. (2016) have suggested that the composition of microalgal cell walls (cellulose, hemicellulose, pectin, and glycoprotein) is a critical issue in enzymes selection to achieve maximal biomass solubilization [175]. As for fermentation, the conversion of microalgal carbohydrates to fermentable sugar is accomplished during the pretreatment process. The combination of sulfuric acid treatment and enzymatic hydrolysis achieves maximal sugar release from Chlorella sorokiniana and Nannochloropsis gaditana (128 mg/g-DCW), while sole sulfuric acid treatment is the most effective method for Scenedesmus almeriensis (88 mg/g-DCW) [176]. Shokrkar et al. (2017) have reported that the combination of three thermostable enzymes achieves effective carbohydrates hydrolysis using wet biomass, which lowers the cost by 30% by avoiding drying [177]. When using microalgae as feedstock in the ABE process, sequential alkali-acid pretreatment of Chlorella vulgaris removes the proteineous substance in biomass, an inhibitor of ABE fermentation bacteria that increases biobutanol yield [171]. Cheng et al. (2017) have showed that hydrolysis and fermentation can be performed simultaneously in a single reactor to lower the total capital cost and minimize contamination risk [178].
The lipid content in microalgal biomass can be extracted and converted to fatty acid methyl esters (FAMEs), major constituents of biodiesel, by an esterification reaction using an acid/base catalysis within supercritical fluid [149,179]. Lipid extraction, which consumes 90% of overall energy, is the bottleneck toward commercialization of microalgal biodiesel. Recent research focuses on developing green solvent and a novel extraction process to minimize the usage of a conventional chloroform-based solvent, which is costly, flammable, and toxic [180]. Pretreating wet/dry microalgal biomass with surfactants effectively ruptures the cell of microalgae, which enables efficient extraction of lipids using a chloroform-free hexane/isopropanol solvent [181]. Orr et al. (2016) have screened over 30 varieties of room-temperature ionic liquids for microalgal biomass rupture, followed by hexane extraction [182]. Among them, [C2mim][EtSO4] gives satisfactory oil recovery of 30 wt% and can be operated with wet microalgal biomass with no deterioration after 5 cycles [182]. The protein and carbohydrate contents in biomass residues after bio-diesel production are richer, which is advantageous to the performance of thermochemical and biological biofuel production [183]. El-Dalatony et al. (2019) have proposed that sequential fermentation of microalgal biomass to bio-ethanol and higher alcohols, followed by biodiesel production from fermentation residue can utilize 89% of biomass [184]. Other than biofuels, valuable biochemicals, such as lutein and C-phycocyanin, can be extracted from microalgae as well [185,186]. To make the process economically feasible, the microalgae farming for CO2 sequestration must be optimized with respect to the fixation rate of CO2, the productivity of valuable biochemical compounds, and the biorefinery process of biomass to biofuels.

4.5. The Circular Bioeconomy

Last but not the least; the awareness of the circular bioeconomy has become an irreplaceable consideration in biotic CO2R. The ultimate goal of the circular bioeconomy is to keep the value of products, materials, and resources as long as possible while minimizing waste generation [187]. As discussed above, the operation of separation and thermal units of biorefinery is energy intensive, which means that achieving a neutral carbon footprint while extracting chemicals via biotic CO2R at the same time remains a technical challenge today. In other words, producing value-added chemicals in biotic CO2R to be economically competitive is equally important to reducing atmospheric CO2 concentration. A recent strengths, weaknesses, opportunities, and threats (SWOT) analysis of organic waste management based on the circular bioeconomy points out that the logistic cost and supply chain management, seasonality, availability of homogenized organic residue, quality and efficiency of the alternative product, and lack of technical standards and regulation are the main threats and weaknesses [188]. The weakness in the availability of homogenized organic residue is particularly true when extracting high-value chemicals such as hydrocarbons and polysaccharides from algae residues. In this case, advances in fermentation technology will be the solution as it narrows the list of fermentation products [189].
Edging conversion technologies that are capable of efficiently destructing the cell walls of algae such as thermocatalytic processes, mechanochemical depolymerization, and their combination will further facilitate the recovery of high-value chemicals [190]. High-energy efficiency of edging conversion technologies will be the additional bonus that further reduces the carbon footprint in biotic CO2R. Of equal importance is the customized biorefinery procedure through optimizing an algal farming strategy and integrating harvest, extraction, fractionation, and purification processes based on the algae cultivation conditions [191]. The above engineering and technology considerations are essential to reaching the goal of value-added products, materials, and the maintenance of resources in the economy toward circular bioeconomy [192]. Apart from individual technology, integrated approaches of microalgae are promising too. For example, integration of microalgae cultivation with seaweed’s anaerobic digestion allows the acquisition of additional biogas. In this case, the methane yield is noted to vary widely depending on the species of microalgae/seaweed and digestion conditions [193]. In-situ transesterification of Chlorella pyrenoidosa with spent coffee grounds enhances a 79.9% higher biodiesel yield and better quality of biodiesel than without spent coffee grounds [194]. The application of waste glycerol, which is the byproduct of biodiesel production from scum-derived oil, is also proposed to balance the C/N ratio of the wastewater in microalgae cultivation [195]. Similarly, the addition of lipid-free algal biomass and waste glycerol in the growth medium for microalgae Scenedesmus obliquus cultivation increases biodiesel production by ~15% [196]. The above integrated approaches will definitely promote the commercialization of market-driven microalgae-derived biofuel production.
In addition to engineering and technology considerations, job creation and new investment opportunities as the result of the development of a new business model is the unique strength of biotic CO2R [188]. For example, based on the circular bioeconomy concept, algae-based feed supplements and recombinant therapeutic production exhibit the potential in response to increasing aquacultural fish demands worldwide [197]. This is because the algal biomass is rich in protein and lipids that offer enormous opportunities to improve animal health, disease resistance, and yields. In fact, being another important sector of the circular bioeconomy, sustainable aquaculture also seeks a condition to reduce waste production while maintaining the value of products, materials, and resources. In this case, algae can be regarded as a valuable source of key nutrients for high quality fish feeds. Recycling algal residues to produce valuable organic matter for further application in fertilizers or amendments is also possible [198]. This means that the nutrients recovered from algae can be returned to nature to further fertilize the stock feed for livestock. Similarly, manures produced by livestock can be utilized for algae farming [199]. As shown in Figure 6, biotic CO2R plays a contributory role to economic and environmental sustainability and is an indispensable link of the circular bioeconomy.

5. Conclusions

The reduction of atmospheric CO2 (CO2R) can be achieved through the abiotic or biotic route. Knowledge of the rational design of efficient abiotic CO2R catalysts and optimization of the biotic CO2R unit process straightforwardly govern the success of CO2R. Advances in catalyst engineering enable delicate control of the surface chemistry of catalysts for adsorption/absorption of CO2/carbonate, surface activation of hydrogen, and stabilization of key intermediates. These surface phenomena not only regulate the abiotic CO2R efficiency but also the selectivity. The latter is another important issue of concern as separation and purification processes require input of external energy. In general, the electrochemical conversion of carbonate in the aqueous phase gives rise to the selective formation of formic acid, while CO2 hydrogenation in the gaseous phase results in the production of CO, CH4, methanol, and hydrocarbons. Further technology breakthroughs are needed to design durable and affordable (electrochemical) catalysts with exceptional CO2 conversion efficiency and selectivity. Promoting biotic CO2R efficiency strongly relies on the resilient microalgae strain and cultivation strategy. The economic feasibility of biotic CO2R remains challenging and additional revenue from microalgae biomass ought to be further maximized through biorefinery by optimizing the cultivation strategy. The latter would be the foundation for the successful implementation of a circular bioeconomy. In addition to the scientific and technological challenges, transdisciplinary cooperation such as the engagement of stakeholders would allow CO2R to be more sustainable.

Author Contributions

Conceptualization, J.-S.C., H.K. and C.-P.H.; methodology, T.W. and J.-Y.L.; writing—original draft preparation, T.W., J.-Y.L., J.-S.C. and C.-M.H.; writing—review and editing, T.W., J.-Y.L., J.-S.C. and C.-M.H.; supervision, C.-D.D. and C.-W.C.; project administration, C.-D.D. and C.-W.C.; funding acquisition, C.-D.D. and C.-W.C. All authors have read and agreed to the published version of the manuscript.

Funding

Much appreciation goes to the Ministry of Science and Technology (MOST) of Taiwan (ROC), under grant number MOST 107-2113-M-007-005-MY2 and 106-2221-E-992-302-MY3, for the financial support. Additional support was provided by the U.S. National Science Foundation through trough NSF grant IOA 1632899 to CPH. H.K. was supported by the Basic Study and Interdisciplinary R&D Foundation Fund of the University of Seoul (2020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Redfield, A.C.; Ketchum, B.K.; Richards, F.A. The influence of organisms on the composition of sea-water. In The Sea; Wiley: New York, NY, USA, 1963; Volume 2, pp. 26–77. [Google Scholar]
  2. Francke, R.; Schille, B.; Roemelt, M. Homogeneously catalyzed electroreduction of carbon dioxide—Methods, mechanisms, and catalysts. Chem. Rev. 2018, 118, 4631–4701. [Google Scholar] [CrossRef] [PubMed]
  3. Qiao, J.; Liu, Y.; Hong, F.; Zhang, J. A review of catalysts for the electroreduction of carbon dioxide to produce low-carbon fuels. Chem. Soc. Rev. 2014, 43, 631–675. [Google Scholar] [CrossRef] [PubMed]
  4. Rakowski Dubois, M.; Dubois, D.L. Development of molecular electrocatalysts for CO2 reduction and H2 production/oxidation. Acc. Chem. Res. 2009, 42, 1974–1982. [Google Scholar] [CrossRef] [PubMed]
  5. Mellmann, D.; Sponholz, P.; Junge, H.; Beller, M. Formic acid as a hydrogen storage material—Development of homogeneous catalysts for selective hydrogen release. Chem. Soc. Rev. 2016, 45, 3954–3988. [Google Scholar] [CrossRef]
  6. Sordakis, K.; Tang, C.; Vogt, L.K.; Junge, H.; Dyson, P.J.; Beller, M.; Laurenczy, G. Homogeneous catalysis for sustainable hydrogen storage in formic acid and alcohols. Chem. Rev. 2018, 118, 372–433. [Google Scholar] [CrossRef] [PubMed]
  7. Huang, J.Z.; Wang, L.S. Bifunctional wood for electrocatalytic CO2 reduction to formate and electroanalytical detection of myricetin and cadmium (II). Electrochim. Acta 2019, 319, 569–576. [Google Scholar] [CrossRef]
  8. Kim, Y.E.; Lee, W.; Youn, M.H.; Jeong, S.K.; Kim, H.J.; Park, J.C.; Park, K.T. Leaching-resistant SnO2/gamma-Al2O3 nanocatalyst for stable electrochemical CO2 reduction into formate. J. Indust. Eng. Chem. 2019, 78, 73–78. [Google Scholar] [CrossRef]
  9. Asset, T.; Garcia, S.T.; Herrera, S.; Andersen, N.; Chen, Y.C.; Peterson, E.J.; Matanovic, I.; Artyushkova, K.; Lee, J.; Minteer, S.D.; et al. Investigating the nature of the active sites for the CO2 reduction reaction on carbon-based electrocatalysts. ACS Catal. 2019, 9, 7668–7678. [Google Scholar] [CrossRef]
  10. Zhang, N.; Gao, C.; Xiong, Y.J. Defect engineering: A versatile tool for tuning the activation of key molecules in photocatalytic reactions. J. Energy Chem. 2019, 37, 43–57. [Google Scholar] [CrossRef] [Green Version]
  11. Zhou, L.Q.; Ling, C.; Zhou, H.; Wang, X.; Liao, J.; Reddy, G.K.; Deng, L.Z.; Peck, T.C.; Zhang, R.G.; Whittingham, M.S.; et al. A high-performance oxygen evolution catalyst in neutral-pH for sunlight-driven CO2 reduction. Nat. Comm. 2019, 10, 4081. [Google Scholar] [CrossRef] [Green Version]
  12. Yang, D.R.; Yu, H.D.; He, T.; Zuo, S.W.; Liu, X.Z.; Yang, H.Z.; Ni, B.; Li, H.Y.; Gu, L.; Wang, D. Visible-light-switched electron transfer over single porphyrin-metal atom center for highly selective electroreduction of carbon dioxide. Nat. Comm. 2019, 10, 3844. [Google Scholar] [CrossRef] [Green Version]
  13. Chang, X.X.; Wang, T.; Zhang, P.; Wei, Y.J.; Zhao, J.B.; Gong, J.L. Stable aqueous photoelectrochemical CO2 reduction by a Cu2O dark cathode with improved selectivity for carbonaceous products. Angew. Chim. Int. Ed. 2016, 55, 8840–8845. [Google Scholar] [CrossRef]
  14. Cheng, J.; Zhang, M.; Liu, J.Z.; Zhou, J.H.; Cen, K.F. A Cu foam cathode used as a Pt-RGO catalyst matrix to improve CO2 reduction in a photoelectrocatalytic cell with a TiO2 photoanode. J. Mater. Chem. A 2015, 3, 12947–12957. [Google Scholar] [CrossRef]
  15. Pletcher, D. The cathodic reduction of carbon dioxide—What can it realistically achieve? A mini review. Electrochem. Comm. 2015, 61, 97–101. [Google Scholar] [CrossRef] [Green Version]
  16. Honoré, M.N.; Belmonte-Ureña, L.J.; Navarro-Velasco, A.; Camacho-Ferre, F. Profit analysis of papaya crops under greenhouses as an alternative to traditional intensive horticulture in southeast Spain. Int. J. Environ. Res. Public Health 2019, 16, 2908. [Google Scholar] [CrossRef] [Green Version]
  17. Hamnett, A. Mechanism and electrocatalysis in the direct methanol fuel cell. Catal. Today 1997, 38, 445–457. [Google Scholar] [CrossRef]
  18. Liu, H.; Song, C.; Zhang, L.; Zhang, J.; Wang, H.; Wilkinson, D.P. A review of anode catalysis in the direct methanol fuel cell. J. Power Sources 2006, 155, 95–110. [Google Scholar] [CrossRef]
  19. Jiang, J.; Wieckowski, A. Prospective direct formate fuel cell. Electrochem. Comm. 2012, 18, 41–43. [Google Scholar] [CrossRef]
  20. Bartrom, A.M.; Haan, J.L. The direct formate fuel cell with an alkaline anion exchange membrane. J. Power Sources 2012, 214, 68–74. [Google Scholar] [CrossRef]
  21. An, L.; Chen, R. Direct formate fuel cells: A review. J. Power Sources 2016, 320, 127–139. [Google Scholar] [CrossRef]
  22. Aresta, M. Carbon Dioxide as Chemical Feedstock; Wiley-VCH: Weinheim, Germany, 2010. [Google Scholar]
  23. Brennführer, A.; Neumann, H.; Beller, M. Palladium-catalyzed carbonylation reactions of aryl halides and related compounds. Angew. Chem. Int. Ed. 2009, 48, 4114–4133. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, Q.; Zhang, H.; Lei, A. Oxidative carbonylation reactions: Organometallic compounds (R-M) or hydrocarbons (R-H) as nucleophiles. Angew. Chem. Int. Ed. 2011, 50, 10788–10799. [Google Scholar] [CrossRef]
  25. Pegis, M.L.; Roberts, J.A.S.; Wasylenko, D.J.; Mader, E.A.; Appel, A.M.; Mayer, J.M. Standard reduction potentials for oxygen and carbon dioxide couples in acetonitrile and N,N-dimethylformamide. Inorg. Chem. 2015, 54, 11883–11888. [Google Scholar] [CrossRef] [PubMed]
  26. Gai, S.L.; Yu, J.L.; Yu, H.; Eagle, J.; Zhao, H.; Lucas, J.; Doroodchi, E.; Moghtaderi, B. Process simulation of a near-zero-carbon-emission power plant using CO2 as the renewable energy storage medium. Int. J. Greenh. Gas Control 2016, 447, 240–249. [Google Scholar] [CrossRef]
  27. Sun, Z.; Zeng, L.; Russell, C.K.; Assabumrungrat, S.; Chen, S.Y.; Duan, L.B.; Xiang, W.G.; Gong, J.L. Solar–wind–bio ecosystem for biomass cascade utilization with multigeneration of formic acid, hydrogen, and graphene. ACS Sustain. Chem. Eng. 2019, 7, 22558–22568. [Google Scholar] [CrossRef]
  28. Zouaoui, N.; Ossonon, B.D.; Fan, M.Y.; Mayilukila, D.; Garbarino, S.; de Silveira, G.; Botton, G.A.; Guay, D.; Tavares, A.C. Electroreduction of CO2 to formate on amine modified Pb electrodes. J. Mater. Chem. A 2019, 7, 11272–11281. [Google Scholar] [CrossRef]
  29. Benson, E.E.; Kubiak, C.P.; Sathrum, A.J.; Smieja, J.M. Electrocatalytic and homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev. 2009, 38, 89–99. [Google Scholar] [CrossRef]
  30. Xu, S.Z.; Carter, E.A. Theoretical insights into heterogeneous (photo)electrochemical CO2 reduction. Chem. Rev. 2019, 119, 6631–6669. [Google Scholar] [CrossRef]
  31. Nitopi, S.; Bertheussen, E.; Scott, S.B.; Liu, X.Y.; 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]
  32. Ma, M.; Trzesniewski, B.J.; Xie, J.; Smith, W.A. Selective and efficient reduction of carbon dioxide to carbon monoxide on oxide-derived nanostructured silver electrocatalysts. Angew. Chem. Int. Ed. 2016, 55, 9748–9752. [Google Scholar] [CrossRef]
  33. Taheri, A.; Thompson, E.J.; Fettinger, J.C.; Berben, L.A. An iron electrocatalyst for selective reduction of CO2 to formate in water: Including thermochemical insights. ACS Catal. 2015, 5, 7140–7151. [Google Scholar] [CrossRef]
  34. Rosas-Hernández, A.; Junge, H.; Beller, M.; Roemelt, M.; Francke, R. Cyclopentadienone iron complexes as efficient and selective catalysts for the electroreduction of CO2 to CO. Catal. Sci. Technol. 2017, 7, 459–465. [Google Scholar] [CrossRef]
  35. Buckley, A.K.; Lee, M.; Cheng, T.; Kazantsev, R.V.; Larson, D.M.; Goddard, W.A., III; Toste, F.D.; Toma, F.M. Electrocatalysis at organic-metal interfaces: Identification of structure-reactivity relationships for CO2 reduction at modified Cu surfaces. J. Am. Chem. Soc. 2019, 141, 7355–7364. [Google Scholar] [CrossRef] [Green Version]
  36. Lobaccaro, P.; Singh, M.R.; Clark, E.L.; Kwon, Y.; Bell, A.T.; Ager, J.W. Effects of temperature and gas-liquid mass transfer on the operation of small electrochemical cells for the quantitative evaluation of CO2 reduction electrocatalysts. Phys. Chem. Chem. Phys. 2016, 18, 26777–26785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Tomisaki, M.; Kasahara, S.; Natsui, K.; Ikemiya, N.; Einaga, Y. Switchable product selectivity in the electrochemical reduction of carbon dioxide using boron-doped diamond electrodes. J. Am. Chem. Soc. 2019, 141, 7414–7420. [Google Scholar] [CrossRef] [PubMed]
  38. Hori, Y. Modern Aspects of Electrochemistry; Vayenas, C.G., White, R.E., Gamboa-Aldeco, M.E., Eds.; Springer: New York, NY, USA, 2008. [Google Scholar]
  39. Taheri, A.; Berben, L.A. Tailoring electrocatalysts for selective CO2 or H+ reduction: Iron carbonyl clusters as a case study. Inorg. Chem. 2016, 55, 378–385. [Google Scholar] [CrossRef] [PubMed]
  40. Waldie, K.M.; Ostericher, A.L.; Reineke, M.H.; Sasayama, A.F.; Kubiak, C.P. Hydricity of transition-metal hydrides: Thermodynamic considerations for CO2 reduction. ACS Catal. 2018, 8, 1313–1324. [Google Scholar] [CrossRef] [Green Version]
  41. Kortlever, R.; Shen, J.; Schouten, K.J.P.; Calle-Vallejo, F.; Koper, M.T.M. Catalysts and reaction pathways for the electrochemical reduction of carbon dioxide. J. Phys. Chem. Lett. 2015, 6, 4073–4082. [Google Scholar] [CrossRef]
  42. Yu, X.; Pickup, P.G. Recent advances in direct formic acid fuel cells (DFAFC). J. Power Sources 2008, 182, 124–132. [Google Scholar] [CrossRef]
  43. Feaster, J.T.; Shi, C.; Cave, E.R.; Hatsukade, T.; Abram, D.N.; Kuhl, K.P.; Hahn, C.; Nørskov, J.K.; 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]
  44. Shinagawa, T.; Larrazábal, G.O.; Martín, A.J.; Krumeich, F.; Pérez-Ramírez, J. Sulfur-modified copper catalysts for the electrochemical reduction of carbon dioxide to formate. ACS Catal. 2018, 8, 837–844. [Google Scholar] [CrossRef] [Green Version]
  45. Huan, T.N.; Simon, P.; Rousse, G.; Génois, I.; Artero, V.; Fontecave, M. Porous dendritic copper: An electrocatalyst for highly selective CO2 reduction to formate in water/ionic liquid electrolyte. Chem. Sci. 2017, 8, 742–747. [Google Scholar] [CrossRef] [Green Version]
  46. Li, C.W.; Ciston, J.; Kanan, M.W. Electroreduction of carbon monoxide to liquid fuel on oxide-derived nanocrystalline copper. Nature 2014, 508, 504–507. [Google Scholar] [CrossRef] [PubMed]
  47. Kuhl, K.P.; Cave, E.R.; Abram, D.N.; Jaramillo, T.F. New insights into the electrochemical reduction of carbon dioxide on metallic copper surfaces. Energy Environ. Sci. 2012, 5, 7050–7059. [Google Scholar] [CrossRef]
  48. Hori, Y.; Murata, A.; Takahashi, R. Formation of hydrocarbons in the electrochemical reduction of carbon dioxide at a copper electrode in aqueous solution. J. Chem. Soc. Faraday Trans. 1989, 1, 2309–2326. [Google Scholar] [CrossRef]
  49. Hori, Y.; Takahashi, R.; Yoshinami, Y.; Murata, A. Electrochemical reduction of CO at a copper electrode. J. Phys. Chem. B 1997, 101, 7075–7081. [Google Scholar] [CrossRef]
  50. Hori, Y.; Murata, A.; Takahashi, R.; Suzuki, S. Electroreduction of CO to CH4 and C2H4 at a copper electrode in aqueous solutions at ambient temperature and pressure. J. Am. Chem. Soc. 1987, 109, 5022–5023. [Google Scholar] [CrossRef]
  51. Li, C.W.; Kanan, M.W. CO2 reduction at low overpotential on Cu electrodes resulting from the reduction of thick Cu2O films. J. Am. Chem. Soc. 2012, 134, 7231–7234. [Google Scholar] [CrossRef] [PubMed]
  52. Gong, M.; Cao, Z.; Liu, W.; Nichols, E.M.; Smith, P.T.; Derrick, J.S.; Liu, Y.S.; Liu, J.; Wen, X.; Chang, C.J. Supramolecular porphyrin cages assembled at molecular–materials interfaces for electrocatalytic co reduction. ACS Cent. Sci. 2017, 3, 1032–1040. [Google Scholar] [CrossRef] [Green Version]
  53. Bertheussen, E.; Verdaguer-Casadevall, A.; Ravasio, D.; Montoya, J.H.; Trimarco, D.B.; Roy, C.; Meier, S.; Wendland, J.; Nørskov, J.K.; Stephens, I.E.L.; et al. Acetaldehyde as an intermediate in the electroreduction of carbon monoxide to ethanol on oxide-derived copper. Angew. Chem. Int. Ed. 2016, 55, 1450–1454. [Google Scholar] [CrossRef] [Green Version]
  54. Zhao, Z.; Zhang, Y.; Li, Y.; Zhao, H.; Quan, X. Electrochemical reduction of carbondioxide to formate with Fe-C electrodes in anaerobic sludge digestion process. Water Res. 2016, 106, 339–343. [Google Scholar] [CrossRef]
  55. Lu, M.; Li, Q.; Liu, J.; Zhang, F.M.; Zhang, L.; Wang, J.L.; Kang, Z.H.; Lan, Y.Q. Installing earth-abundant metal active centers to covalent organic frameworks for efficient heterogeneous photocatalytic CO2 reduction. Appl. Catal. B Environ. 2019, 254, 624–633. [Google Scholar] [CrossRef]
  56. Swierk, J.R.; Mallouk, T.E. Design and development of photoanodes for water-splitting dye-sensitized photoelectrochemical cells. Chem. Soc. Rev. 2013, 42, 2357–2387. [Google Scholar] [CrossRef] [PubMed]
  57. Todorova, T.K.; Huan, T.N.; Wang, X.; Agarwala, H.; Fontecave, M. Controlling hydrogen evolution during photoreduction of CO2 to formic acid using [Rh(R-bpy)(Cp*)Cl](+) catalysts: A structure-activity study. Inorg. Chem. 2019, 58, 6893–6903. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Chardon-Noblat, S.; Deronzier, A.; Ziessel, R.; Zsoldos, D. Electroreduction of CO2 catalyzed by polymeric [Ru(bpy)(CO)2]n films in aqueous media: Parameters influencing the reaction selectivity. J. Electroanal. Chem. 1998, 444, 253–260. [Google Scholar] [CrossRef]
  59. Chen, J.; Yin, J.; Zheng, X.P.; Ahsaine, H.A.; Zhou, Y.; Dong, C.W.; Mohammed, O.F.; Takanabe, K.; Bakr, O.M. Compositionally screened eutectic catalytic coatings on halide perovskite photocathodes for photoassisted selective CO2 reduction. ACS Energy Lett. 2019, 4, 1279–1286. [Google Scholar] [CrossRef] [Green Version]
  60. Ajmal, S.; Yang, Y.; Li, K.J.; Tahir, M.A.; Liu, Y.Y.; Wang, T.; Bacha, A.U.R.; Feng, Y.Q.; Deng, Y.; Zhang, L.W. Zinc-modified copper catalyst for efficient (photo-) electrochemical CO2 reduction with high selectivity of HCOOH production. J. Phys. Chem. C 2019, 123, 11555–11563. [Google Scholar] [CrossRef]
  61. An, X.W.; Li, S.S.; Yoshida, A.; Wang, Z.D.; Hao, X.G.; Abudula, A.; Guan, G.Q. Electrodeposition of tin-based electrocatalysts with different surface tin species distributions for electrochemical reduction of CO2 to HCOOH. ACS Sustain. Chem. Eng. 2019, 7, 9360–9368. [Google Scholar] [CrossRef]
  62. Chatterjee, S.; Griego, C.; Hart, J.L.; Li, Y.W.; Taheri, M.L.; Keith, J.; Snyder, J.D. Free standing nanoporous palladium alloys as CO poisoning tolerant electrocatalysts for the electrochemical reduction of CO2 to formate. ACS Catal. 2019, 6, 5290–5301. [Google Scholar] [CrossRef] [Green Version]
  63. Raju, R.K.; Rodriguez, P.; Johnston, R.L. Can a single valence electron alter the electrocatalytic activity and selectivity for CO2 reduction on the subnanometer scale? J. Phys. Chem. C 2019, 123, 14591–14609. [Google Scholar] [CrossRef]
  64. Qin, C.L.; Fan, A.X.; Zhang, X.; Dai, X.P.; Sun, H.; Ren, D.H.; Dong, Z.; Wang, Y.; Luan, C.L.; Ye, J.Y.; et al. The in situ etching assisted synthesis of Pt-Fe-Mn ternary alloys with high-index facets as efficient catalysts for electro-oxidation reactions. Nanoscale 2019, 11, 9061–9075. [Google Scholar] [CrossRef]
  65. Pan, F.P.; Li, B.Y.; Deng, W.; Du, Z.C.; Gang, Y.; Wang, G.F.; Li, Y. Promoting electrocatalytic CO2 reduction on nitrogen-doped carbon with sulfur addition. Appl. Catal. B Environ. 2019, 252, 240–249. [Google Scholar] [CrossRef]
  66. Yin, Y.; Gao, D.F.; Yao, S.Y.; Zhao, B.; Cai, F.; Lin, L.L.; Tang, P.; Zhai, P.; Wang, G.X.; Ma, D.; et al. Highly selective palladium-copper bimetallic electrocatalysts for the electrochemical reduction of CO2 to CO. Nano Energy 2016, 27, 35–43. [Google Scholar] [CrossRef]
  67. Gong, Q.F.; Ding, P.; Xu, M.Q.; Zhu, X.R.; Wang, M.Y.; Deng, J.; Ma, Q.; Han, N.; Zhu, Y.; Lu, J.; et al. Structural defects on converted bismuth oxide nanotubes enable highly active electrocatalysis of carbon dioxide reduction. Nat. Comm. 2019, 10, 2807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Gao, D.; Zhou, H.; Wang, J.; Miao, S.; Yang, F.; Wang, G.; Wang, J.; Bao, X. Size-dependent electrocatalytic reduction of CO2 over Pd nanoparticles. J. Am. Chem. Soc. 2015, 137, 4288–4291. [Google Scholar] [CrossRef]
  69. Jiang, B.; Zhang, X.G.; Jiang, K.; Wu, D.Y.; Cai, W.B. Boosting formate production in electrocatalytic CO2 reduction over wide potential window on Pd surfaces. J. Am. Chem. Soc. 2018, 140, 2880–2889. [Google Scholar] [CrossRef] [PubMed]
  70. Eder, G.M.; Pyles, D.A.; Wolfson, E.R.; McGrier, P.L. A ruthenium porphyrin-based porous organic polymer for the hydrosilylative reduction of CO2 to formate. Chem. Comm. 2019, 55, 7195–7198. [Google Scholar] [CrossRef]
  71. Weetman, C.; Bag, P.; Szilvasi, T.; Jandl, C.; Inoue, S. CO2 Fixation and catalytic reduction by a neutral aluminum double bond. Angew. Chem. Int. Ed. 2019, 58, 10961–10965. [Google Scholar] [CrossRef] [Green Version]
  72. Zhao, Y.; Wang, T.H.; Wang, Y.B.; Hao, R.L.; Wang, H.; Han, Y.H. Catalytic reduction of CO2 to HCO2 by nanoscale nickel-based bimetallic alloy under atmospheric pressure. J. Ind. Eng. Chem. 2019, 77, 291–302. [Google Scholar] [CrossRef]
  73. Porosoff, M.D.; Yan, B.; Chen, J.G. Catalytic reduction of CO2 by H2 for synthesis of CO, methanol and hydrocarbons: Challenges and opportunities. Energy Environ. Sci. 2016, 9, 62–73. [Google Scholar] [CrossRef]
  74. Rodriguez, J.A.; Liu, P.; Stacchiola, D.J.; Senanayake, S.D.; White, M.G.; Chen, J.G. Hydrogenation of CO2 to methanol: Importance of metal-oxide and metal-carbide interfaces in the activation of CO2. ACS Catal. 2015, 5, 6696–6706. [Google Scholar] [CrossRef]
  75. Gao, P.; Dang, S.; Li, S.; Bu, X.; Liu, Z.; Qiu, M.; Yang, C.; Wang, H.; Zhong, L.; Han, Y.; et al. Direct production of lower olefins from CO2 conversion via bifunctional catalysis. ACS Catal. 2018, 8, 571–578. [Google Scholar] [CrossRef]
  76. Sun, K.; Fan, Z.; Ye, J.; Yan, J.; Ge, Q.; Li, Y.; He, W.; Yang, W.; Liu, C. Hydrogenation of CO2 to methanol over In2O3 catalyst. J. CO2 Util. 2015, 12, 1–6. [Google Scholar] [CrossRef]
  77. Gnanamani, M.K.; Jacobs, G.; Hamdeh, H.H.; Shafer, W.D.; Liu, F.; Hopps, S.D.; Thomas, G.A.; Davis, B.H. Hydrogenation of carbon dioxide over Co-Fe bimetallic catalysts. ACS Catal. 2016, 6, 913–927. [Google Scholar] [CrossRef]
  78. Dreyer, J.A.H.; Li, P.; Zhang, L.; Beh, G.K.; Zhang, R.; Sit, P.H.L.; Teoh, W.Y. Influence of the oxide support reducibility on the CO2 methanation over Ru-based catalysts. Appl. Catal. B Environ. 2017, 219, 715–726. [Google Scholar] [CrossRef]
  79. Li, W.; Nie, X.; Jiang, X.; Zhang, A.; Ding, F.; Liu, M.; Liu, Z.; Guo, X.; Song, C. ZrO2 support imparts superior activity and stability of Co catalysts for CO2 methanation. Appl. Catal. B Environ. 2018, 220, 397–408. [Google Scholar] [CrossRef]
  80. Zhan, G.; Zeng, H.C. ZIF-67-derived nanoreactors for controlling product selectivity in CO2 hydrogenation. ACS Catal. 2017, 7, 7509–7519. [Google Scholar] [CrossRef]
  81. Li, Y.; Cai, X.; Chen, S.; Zhang, H.; Zhang, K.H.L.; Hong, J.; Chen, B.; Kuo, D.H.; Wang, W. Highly dispersed metal carbide on ZIF-derived pyridinic-N-doped carbon for CO2 enrichment and selective hydrogenation. ChemSusChem 2018, 11, 1040–1047. [Google Scholar] [CrossRef] [PubMed]
  82. Kattel, S.; Yu, W.; Yang, X.; Yan, B.; Huang, Y.; Wan, W.; Liu, P.; Chen, J.G. CO2 hydrogenation over oxide-supported PtCo catalysts: The role of the oxide support in determining the product selectivity. Angew. Chem. Int. Ed. 2016, 55, 7968–7973. [Google Scholar] [CrossRef]
  83. Ren, J.; Qin, X.; Yang, J.Z.; Qin, Z.F.; Guo, H.L.; Lin, J.Y.; Li, Z. Methanation of carbon dioxide over Ni-M/ZrO2 (M = Fe, Co, Cu) catalysts: Effect of addition of a second metal. Fuel Process. Technol. 2015, 137, 204–211. [Google Scholar] [CrossRef]
  84. Kirchner, J.; Anolleck, J.K.; Loesch, H.; Kureti, S. Methanation of CO2 on iron based catalysts. Appl. Catal. B Environ. 2018, 223, 47–59. [Google Scholar] [CrossRef]
  85. Bi, Q.; Huang, X.; Yin, G.; Chen, T.; Du, X.; Cai, J.; Xu, J.; Liu, Z.; Han, Y.; Huang, F. Cooperative catalysis of nickel and nickel oxide for efficient reduction of CO2 to CH4. ChemCatChem 2019, 11, 1295–1302. [Google Scholar] [CrossRef]
  86. Dietz, L.; Piccinin, S.; Maestri, M. Mechanistic insights into CO2 activation via reverse water-gas shift on metal surfaces. J. Phys. Chem. C 2015, 119, 4959–4966. [Google Scholar] [CrossRef]
  87. Kattel, S.; Yan, B.; Chen, J.G.; Liu, P. CO2 hydrogenation on Pt, Pt/SiO2 and Pt/TiO2: Importance of synergy between Pt and oxide support. J. Catal. 2016, 343, 115–126. [Google Scholar] [CrossRef] [Green Version]
  88. Yang, X.; Su, X.; Chen, X.; Duan, H.; Liang, B.; Liu, Q.; Liu, X.; Ren, Y.; Huang, Y.; Zhang, T. Promotion effects of potassium on the activity and selectivity of Pt/zeolite catalysts for reverse water gas shift reaction. Appl. Catal. B Environ. 2017, 216, 95–105. [Google Scholar] [CrossRef]
  89. Wang, X.; Shi, H.; Kwak, J.H.; Szanyi, J. Mechanism of CO2 hydrogenation on Pd/Al2O3 catalysts: Kinetics and transient DRIFTS-MS studies. ACS Catal. 2015, 5, 6337–6349. [Google Scholar] [CrossRef]
  90. Guo, Y.; Mei, S.; Yuan, K.; Wang, D.J.; Liu, H.C.; Yan, C.H.; Zhang, Y.W. Low-temperature CO2 methanation over CeO2-supported Ru single atoms, nanoclusters, and nanoparticles competitively tuned by strong metal-support interactions and H-spillover effect. ACS Catal. 2018, 8, 6203–6215. [Google Scholar] [CrossRef]
  91. Xu, J.; Su, X.; Duan, H.; Hou, B.; Lin, Q.; Liu, X.; Pan, X.; Pei, G.; Geng, H.; Huang, Y.; et al. Influence of pretreatment temperature on catalytic performance of rutile TiO2-supported ruthenium catalyst in CO2 methanation. J. Catal. 2016, 333, 227–237. [Google Scholar] [CrossRef]
  92. Wu, H.C.; Chang, Y.C.; Wu, J.H.; Lin, J.H.; Lin, I.K.; Chen, C.S. Methanation of CO2 and reverse water gas shift reactions on Ni/SiO2 catalysts: The influence of particle size on selectivity and reaction pathway. Catal. Sci. Technol. 2015, 5, 4154–4163. [Google Scholar] [CrossRef]
  93. Ren, H.; Xu, C.H.; Zhao, H.Y.; Wang, Y.X.; Liu, J.; Liu, J.Y. Methanol synthesis from CO2 hydrogenation over Cu/gamma-Al2O3 catalysts modified by ZnO, ZrO2 and MgO. J. Ind. Eng. Chem. 2015, 28, 261–267. [Google Scholar] [CrossRef]
  94. Hus, M.; Dasireddy, V.D.B.C.; Stefancic, N.S.; Likozar, B. Mechanism, kinetics and thermodynamics of carbon dioxide hydrogenation to methanol on Cu/ZnAl2O4 spinel-type heterogeneous catalysts. Appl. Catal. B Environ. 2017, 207, 267–278. [Google Scholar] [CrossRef]
  95. Studt, F.; Behrens, M.; Kunkes, E.L.; Thomas, N.; Zander, S.; Tarasov, A.; Schumann, J.; Frei, E.; Varley, J.B.; Abild-Pedersen, F.; et al. The mechanism of CO and CO2 hydrogenation to methanol over Cu-based catalysts. ChemCatChem 2015, 7, 1105–1111. [Google Scholar] [CrossRef] [Green Version]
  96. Kunkes, E.L.; Studt, F.; Abild-Pedersen, F.; Schloegl, R.; Behrens, M. Hydrogenation of CO2 to methanol and CO on Cu/ZnO/Al2O3: Is there a common intermediate or not? J. Catal. 2015, 328, 43–48. [Google Scholar] [CrossRef] [Green Version]
  97. Karelovic, A.; Ruiz, P. The role of copper particle size in low pressure methanol synthesis via CO2 hydrogenation over Cu/ZnO catalysts. Catal. Sci. Technol. 2015, 5, 869–881. [Google Scholar] [CrossRef]
  98. Ro, I.; Liu, Y.; Ball, M.R.; Jackson, D.H.K.; Chada, J.P.; Sener, C.; Kuech, T.F.; Madon, R.J.; Huber, G.W.; Dumesic, J.A. Role of the Cu-ZrO2 interfacial sites for conversion of ethanol to ethyl acetate and synthesis of methanol from CO2 and H2. ACS Catal. 2016, 6, 7040–7050. [Google Scholar] [CrossRef]
  99. Phongamwong, T.; Chantaprasertporn, U.; Witoon, T.; Numpilai, T.; Poo-Arporn, Y.; Limphirat, W.; Donphai, W.; Dittanet, P.; Chareonpanich, M.; Limtrakul, J. CO2 hydrogenation to methanol over CuO-ZnO-ZrO2-SiO2 catalysts: Effects of SiO2 contents. Chem. Eng. J. 2017, 316, 692–703. [Google Scholar] [CrossRef]
  100. Jiang, X.; Koizumi, N.; Guo, X.; Song, C. Bimetallic Pd-Cu catalysts for selective CO2 hydrogenation to methanol. Appl. Catal. B Environ. 2015, 170, 173–185. [Google Scholar] [CrossRef]
  101. Bahruji, H.; Bowker, M.; Hutchings, G.; Dimitratos, N.; Wells, P.; Gibson, E.; Jones, W.; Brookes, C.; Morgan, D.; Lalev, G. Pd/ZnO catalysts for direct CO2 hydrogenation to methanol. J. Catal. 2016, 343, 133–146. [Google Scholar] [CrossRef] [Green Version]
  102. Nie, X.; Jiang, X.; Wang, H.; Luo, W.; Janik, M.J.; Chen, Y.; Guo, X.; Song, C. Mechanistic understanding of alloy effect and water promotion for Pd-Cu bimetallic catalysts in CO2 hydrogenation to methanol. ACS Catal. 2018, 8, 4873–4892. [Google Scholar] [CrossRef]
  103. Stangeland, K.; Kalai, D.Y.; Ding, Y.; Yu, Z. Mesoporous manganese-cobalt oxide spinel catalysts for CO2 hydrogenation to methanol. J. CO2 Util. 2019, 32, 146–154. [Google Scholar] [CrossRef]
  104. Martin, O.; Martín, A.J.; Mondelli, C.; Mitchell, S.; Segawa, T.F.; Hauert, R.; Drouilly, C.; Curulla-Ferré, D.; Pérez-Ramírez, J. Indium oxide as a superior catalyst for methanol synthesis by CO2 Hydrogenation. Angew. Chem. Int. Ed. 2016, 55, 6261–6265. [Google Scholar] [CrossRef]
  105. Wei, J.; Ge, Q.; Yao, R.; Wen, Z.; Fang, C.; Guo, L.; Xu, H.; Sun, J. Directly converting CO2 into a gasoline fuel. Nat. Comm. 2017, 8, 15174. [Google Scholar] [CrossRef] [Green Version]
  106. Visconti, C.G.; Martinelli, M.; Falbo, L.; Infantes-Molina, A.; Lietti, L.; Forzatti, P.; Iaquaniello, G.; Palo, E.; Picutti, B.; Brignoli, F. CO2 hydrogenation to lower olefins on a high surface area K-promoted bulk Fe-catalyst. Appl. Catal. B Environ. 2017, 200, 530–542. [Google Scholar] [CrossRef]
  107. Albrecht, M.; Rodemerck, U.; Schneider, M.; Broering, M.; Baabe, D.; Kondratenko, E.V. Unexpectedly efficient CO2 hydrogenation to higher hydrocarbons over non-doped Fe2O3. Appl. Catal. B Environ. 2017, 204, 119–126. [Google Scholar] [CrossRef]
  108. Liu, J.; Zhang, A.; Liu, M.; Hu, S.; Ding, F.; Song, C.; Guo, X. Fe-MOF-derived highly active catalysts for carbon dioxide hydrogenation to valuable hydrocarbons. J. CO2 Util. 2017, 21, 100–107. [Google Scholar] [CrossRef]
  109. Choi, Y.H.; Jang, Y.J.; Park, H.; Kim, W.Y.; Lee, Y.H.; Choi, S.H.; Lee, J.S. Carbon dioxide Fischer-Tropsch synthesis: A new path to carbon-neutral fuels. Appl. Catal. B Environ. 2017, 202, 605–610. [Google Scholar] [CrossRef]
  110. Wei, J.; Sun, J.; Wen, Z.; Fang, C.; Ge, Q.; Xu, H. New insights into the effect of sodium on Fe3O4-based nanocatalysts for CO2 hydrogenation to light olefins. Catal. Sci. Technol. 2016, 6, 4786–4793. [Google Scholar] [CrossRef]
  111. Satthawong, R.; Koizumi, N.; Song, C.; Prasassarakich, P. Light olefin synthesis from CO2 hydrogenation over K-promoted Fe-Co bimetallic catalysts. Catal. Today 2015, 251, 34–40. [Google Scholar] [CrossRef]
  112. Gao, P.; Li, S.; Bu, X.; Dang, S.; Liu, Z.; Wang, H.; Zhong, L.; Qiu, M.; Yang, C.; Cai, J.; et al. Direct conversion of CO2 into liquid fuels with high selectivity over a bifunctional catalyst. Nat. Chem. 2017, 9, 1019–1024. [Google Scholar] [CrossRef]
  113. Cheah, W.Y.; Show, P.L.; Chang, J.S.; Ling, T.C.; Juan, J.C. Biosequestration of atmospheric CO2 and flue gas-containing CO2 by microalgae. Bioresour. Technol. 2015, 184, 190–201. [Google Scholar] [CrossRef] [PubMed]
  114. Yadav, G.; Sen, R. Microalgal green refinery concept for biosequestration of carbon-dioxide vis-à-vis wastewater remediation and bioenergy production: Recent technological advances in climate research. J. CO2 Util. 2017, 17, 188–206. [Google Scholar] [CrossRef]
  115. Kuo, C.M.; Jian, J.F.; Sun, Y.L.; Lin, T.H.; Yang, Y.C.; Zhang, W.X.; Chang, H.F.; Lai, J.T.; Chang, J.S.; Lin, C.S. An efficient photobioreactors/raceway circulating system combined with alkaline-CO2 capturing medium for microalgal cultivation. Bioresour. Technol. 2018, 266, 398–406. [Google Scholar] [CrossRef] [PubMed]
  116. Duan, Y.; Shi, F. Bioreactor design for algal growth as a sustainable energy source. In Reactor and Process Design in Sustainable Energy Technology; Elsevier: Amsterdam, The Netherlands, 2014; pp. 27–60. [Google Scholar]
  117. Xie, Y.P.; Ho, S.H.; Chen, C.Y.; Chen, C.N.N.; Liu, C.C.; Ng, I.S.; Jing, K.J.; Yang, S.C.; Chen, C.H.; Chang, J.S.; et al. Simultaneous enhancement of CO2 fixation and lutein production with thermo-tolerant Desmodesmus sp. F51 using a repeated fed-batch cultivation strategy. Biochem. Eng. J. 2014, 86, 33–40. [Google Scholar] [CrossRef]
  118. Gonçalves, A.L.; Rodrigues, C.M.; Pires, J.C.M.; Simões, M. The effect of increasing CO2 concentrations on its capture, biomass production and wastewater bioremediation by microalgae and cyanobacteria. Algal Res. 2016, 14, 127–136. [Google Scholar] [CrossRef]
  119. Chen, H.; Wang, J.; Zheng, Y.; Zhan, J.; He, C.; Wang, Q. Algal biofuel production coupled bioremediation of biomass power plant wastes based on Chlorella sp C2 cultivation. Appl. Energy 2018, 211, 296–305. [Google Scholar] [CrossRef]
  120. Jain, D.; Ghonse, S.S.; Trivedi, T.; Fernandes, G.L.; Menezes, L.D.; Damare, S.R.; Mamatha, S.S.; Kumar, S.; Gupta, V. CO2 fixation and production of biodiesel by Chlorella vulgaris NIOCCV under mixotrophic cultivation. Bioresour. Technol. 2019, 273, 672–676. [Google Scholar] [CrossRef]
  121. Kumar, K.; Banerjee, D.; Das, D. Carbon dioxide sequestration from industrial flue gas by Chlorella sorokiniana. Bioresour. Technol. 2014, 152, 225–233. [Google Scholar] [CrossRef]
  122. Tripathi, R.; Singh, J.; Thakur, I.S. Characterization of microalga Scenedesmus sp ISTGA1 for potential CO2 sequestration and biodiesel production. Renew. Energy 2015, 74, 774–781. [Google Scholar] [CrossRef]
  123. Varshney, P.; Beardall, J.; Bhattacharya, S.; Wangikar, P.P. Isolation and biochemical characterisation of two thermophilic green algal species- Asterarcys quadricellulare and Chlorella sorokiniana, which are tolerant to high levels of carbon dioxide and nitric oxide. Algal Res. 2018, 30, 28–37. [Google Scholar] [CrossRef]
  124. Yun, H.S.; Ji, M.K.; Park, Y.T.; Salama, E.S.; Choi, J. Microalga, Acutodesmus obliquus KGE 30 as a potential candidate for CO2 mitigation and biodiesel production. Environ. Sci. Pollut. Res. 2016, 23, 17831–17839. [Google Scholar] [CrossRef]
  125. Bhakta, J.N.; Lahiri, S.; Pittman, J.K.; Jana, B.B. Carbon dioxide sequestration in wastewater by a consortium of elevated carbon dioxide-tolerant microalgae. J. CO2 Util. 2015, 10, 105–112. [Google Scholar] [CrossRef]
  126. Kuo, C.M.; Lin, T.H.; Yang, Y.C.; Zhang, W.X.; Lai, J.T.; Wu, H.T.; Chang, J.S.; Lin, C.S. Ability of an alkali-tolerant mutant strain of the microalga Chlorella sp. AT1 to capture carbon dioxide for increasing carbon dioxide utilization efficiency. Bioresour. Technol. 2017, 244, 243–251. [Google Scholar] [CrossRef] [PubMed]
  127. Yang, B.; Liu, J.; Ma, X.; Guo, B.; Liu, B.; Wu, T.; Jiang, Y.; Chen, F. Genetic engineering of the Calvin cycle toward enhanced photosynthetic CO2 fixation in microalgae. Biotechnol. Biofuels 2017, 10, 229. [Google Scholar] [CrossRef] [PubMed]
  128. Li, D.; Wang, L.; Zhao, Q.; Wei, W.; Sun, Y. Improving high carbon dioxide tolerance and carbon dioxide fixation capability of Chlorella sp by adaptive laboratory evolution. Bioresour. Technol. 2015, 185, 269–275. [Google Scholar] [CrossRef]
  129. Bilanovic, D.; Holland, M.; Starosvetsky, J.; Armon, R. Co-cultivation of microalgae and nitrifiers for higher biomass production and better carbon capture. Bioresour. Technol. 2016, 220, 282–288. [Google Scholar] [CrossRef]
  130. Aslam, A.; Thomas-Hall, S.R.; Mughal, T.A.; Schenk, P.M. Selection and adaptation of microalgae to growth in 100% unfiltered coal-fired flue gas. Bioresour. Technol. 2017, 233, 271–283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Nayak, M.; Suh, W.I.; Lee, B.; Chang, Y.K. Enhanced carbon utilization efficiency and FAME production of Chlorella sp HS2 through combined supplementation of bicarbonate and carbon dioxide. Energy Convers. Manag. 2018, 156, 45–52. [Google Scholar] [CrossRef]
  132. Jiang, Y.; Zhang, W.; Wang, J.; Chen, Y.; Shen, S.; Liu, T. Utilization of simulated flue gas for cultivation of Scenedesmus dimorphus. Bioresour. Technol. 2013, 128, 359–364. [Google Scholar] [CrossRef]
  133. Vo, H.N.P.; Bui, X.T.; Nguyen, T.T.; Nguyen, D.D.; Dao, T.S.; Cao, N.D.T.; Vo, T.K.Q. Effects of nutrient ratios and carbon dioxide bio-sequestration on biomass growth of Chlorella sp in bubble column photobioreactor. J. Environ. Manag. 2018, 219, 1–8. [Google Scholar] [CrossRef]
  134. Liu, X.; Ying, K.; Chen, G.; Zhou, C.; Zhang, W.; Zhang, X.; Cai, Z.; Holmes, T.; Tao, Y. Growth of Chlorella vulgaris and nutrient removal in the wastewater in response to intermittent carbon dioxide. Chemosphere 2017, 186, 977–985. [Google Scholar] [CrossRef]
  135. Posadas, E.; Morales, M.d.M.; Gomez, C.; Acién, F.G.; Muñoz, R. Influence of pH and CO2 source on the performance of microalgae-based secondary domestic wastewater treatment in outdoors pilot raceways. Chem. Eng. J. 2015, 265, 239–248. [Google Scholar] [CrossRef] [Green Version]
  136. Kuo, C.M.; Jian, J.F.; Lin, T.H.; Chang, Y.B.; Wan, X.H.; Lai, J.T.; Chang, J.S.; Lin, C.S. Simultaneous microalgal biomass production and CO2 fixation by cultivating Chlorella sp. GD with aquaculture wastewater and boiler flue gas. Bioresour. Technol. 2016, 221, 241–250. [Google Scholar] [CrossRef]
  137. Eze, V.C.; Velasquez-Orta, S.B.; Hernandez-Garcia, A.; Monje-Ramirez, I.; Orta-Ledesma, M.T. Kinetic modelling of microalgae cultivation for wastewater treatment and carbon dioxide sequestration. Algal Res. 2018, 32, 131–141. [Google Scholar] [CrossRef]
  138. Fu, W.; Gudmundsson, S.; Wichuk, K.; Palsson, S.; Palsson, B.O.; Salehi-Ashtiani, K.; Brynjolfsson, S. Sugar-stimulated CO2 sequestration by the green microalga Chlorella vulgaris. Sci. Total Environ. 2019, 654, 275–283. [Google Scholar] [CrossRef]
  139. Dineshkumar, R.; Subramanian, G.; Dash, S.K.; Sen, R. Development of an optimal light-feeding strategy coupled with semi-continuous reactor operation for simultaneous improvement of microalgal photosynthetic efficiency, lutein production and CO2 sequestration. Biochem. Eng. J. 2016, 113, 47–56. [Google Scholar] [CrossRef]
  140. Jacob-Lopes, E.; Franco, T.T. From oil refinery to microalgal biorefinery. J. CO2 Util. 2013, 2, 1–7. [Google Scholar] [CrossRef]
  141. Kassim, M.A.; Meng, T.K. Carbon dioxide (CO2) biofixation by microalgae and its potential for biorefinery and biofuel production. Sci. Total Environ. 2017, 584, 1121–1129. [Google Scholar] [CrossRef]
  142. Kao, C.Y.; Chen, T.Y.; Chang, Y.B.; Chiu, T.W.; Lin, H.Y.; Chen, C.D.; Chang, J.S.; Lin, C.S. Utilization of carbon dioxide in industrial flue gases for the cultivation of microalga Chlorella sp. Bioresour. Technol. 2014, 166, 485–493. [Google Scholar] [CrossRef] [PubMed]
  143. Huang, Y.; Sun, Y.; Liao, Q.; Fu, Q.; Xia, A.; Zhu, X. Improvement on light penetrability and microalgae biomass production by periodically pre-harvesting Chlorella vulgaris cells with culture medium recycling. Bioresour. Technol. 2016, 216, 669–676. [Google Scholar] [CrossRef]
  144. Moheimani, N.R. Tetraselmis suecica culture for CO2 bioremediation of untreated flue gas from a coal-fired power station. J. Appl. Phycol. 2016, 28, 2139–2146. [Google Scholar] [CrossRef]
  145. Ou-Yang, C.; Chen, H.W.; Ho, C.H.; Chou, J.C.; Yuan, Y.T.; Ho, C.L.; Hsueh, H.T.; Chen, S.T.; Liao, P.C.; Chao, L.K. Value chain analysis of algal bioenergy and carbon capture integrated with a biotechnology innovation. J. Clean. Prod. 2018, 180, 349–359. [Google Scholar] [CrossRef]
  146. Zhu, L.D.; Xu, Z.B.; Qin, L.; Wang, Z.M.; Hiltunen, E.; Li, Z.H. Oil production from pilot-scale microalgae cultivation: An economics evaluation. Energy Sources Part B Econ. Plan. Policy 2016, 11, 11–17. [Google Scholar] [CrossRef]
  147. Rezvani, S.; Moheimani, N.R.; Bahri, P.A. Techno-economic assessment of CO2 bio-fixation using microalgae in connection with three different state-of-the-art power plants. Comput. Chem. Eng. 2016, 84, 290–301. [Google Scholar] [CrossRef] [Green Version]
  148. Venkata Mohan, S.; Modestra, J.A.; Amulya, K.; Butti, S.K.; Velvizhi, G. A circular bioeconomy with biobased products from CO2 sequestration. Trends Biotechnol. 2016, 34, 506–519. [Google Scholar] [CrossRef] [PubMed]
  149. Chew, K.W.; Yap, J.Y.; Show, P.L.; Suan, N.H.; Juan, J.C.; Ling, T.C.; Lee, D.J.; Chang, J.S. Microalgae biorefinery: High value products perspectives. Bioresour. Technol. 2017, 229, 53–62. [Google Scholar] [CrossRef]
  150. Duque-Acevedo, M.; Belmonte-Ureña, L.J.; Plaza-Úbeda, J.A.; Camacho-Ferre, F. The management of agricultural waste biomass in the framework of circular economy and bioeconomy: An opportunity for greenhouse agriculture in southeast Spain. Agronomy 2020, 10, 489. [Google Scholar] [CrossRef] [Green Version]
  151. Dębowski, M.; Zieliński, M.; Kazimierowicz, J.; Kujawska, N.; Talbierz, S. Microalgae cultivation technologies as an opportunity for bioenergetic system development—Advantages and limitations. Sustainability 2020, 12, 9980. [Google Scholar] [CrossRef]
  152. Khan, M.I.; Shin, J.H.; Kim, J.D. The promising future of microalgae: Current status, challenges, and optimization of a sustainable and renewable industry for biofuels, feed, and other products. Microb. Cell Factories 2018, 17, 36. [Google Scholar] [CrossRef]
  153. Alobwede, E.; Leake, J.R.; Pandhal, J. Circular economy fertilization: Testing micro and macro algal species as soil improvers and nutrient sources for crop production in greenhouse and field conditions. Geoderma 2019, 334, 113–123. [Google Scholar] [CrossRef]
  154. Sprague, M.; Dick, J.R.; Tocher, D.R. Impact of sustainable feeds on omega-3 long-chain fatty acid levels in farmed Atlantic salmon, 2006–2015. Sci. Rep. 2016, 6, 21892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Dineshbabu, G.; Goswami, G.; Kumar, R.; Sinha, A.; Das, D. Microalgae—Nutritious, sustainable aqua- and animal feed source. J. Funct. Foods 2019, 62, 103545. [Google Scholar] [CrossRef]
  156. Shakya, R.; Whelen, J.; Adhikari, S.; Mahadevan, R.; Neupane, S. Effect of temperature and Na2CO3 catalyst on hydrothermal liquefaction of algae. Algal Res. 2015, 12, 80–90. [Google Scholar] [CrossRef] [Green Version]
  157. Gai, C.; Zhang, Y.; Chen, W.T.; Zhang, P.; Dong, Y. An investigation of reaction pathways of hydrothermal liquefaction using Chlorella pyrenoidosa and Spirulina platensis. Energy Convers. Manag. 2015, 96, 330–339. [Google Scholar] [CrossRef]
  158. Chen, Y.; Mu, R.; Yang, M.; Fang, L.; Wu, Y.; Wu, K.; Liu, Y.; Gong, J. Catalytic hydrothermal liquefaction for bio-oil production over CNTs supported metal catalysts. Chem. Eng. Sci. 2017, 161, 299–307. [Google Scholar] [CrossRef]
  159. López Barreiro, D.; Bauer, M.; Hornung, U.; Posten, C.; Kruse, A.; Prins, W. Cultivation of microalgae with recovered nutrients after hydrothermal liquefaction. Algal Res. 2015, 9, 99–106. [Google Scholar] [CrossRef]
  160. Chen, W.H.; Lin, B.J.; Huang, M.Y.; Chang, J.S. Thermochemical conversion of microalgal biomass into biofuels: A review. Bioresour. Technol. 2015, 184, 314–327. [Google Scholar] [CrossRef] [PubMed]
  161. Raheem, A.; Dupont, V.; Channa, A.Q.; Zhao, X.; Vuppaladadiyam, A.K.; Taufiq-Yap, Y.H.; Zhao, M.; Harun, R. Parametric characterization of air gasification of Chlorella vulgaris biomass. Energy Fuels 2017, 31, 2959–2969. [Google Scholar] [CrossRef] [Green Version]
  162. Tiong, L.; Komiyama, M.; Uemura, Y.; Nguyen, T.T. Catalytic supercritical water gasification of microalgae: Comparison of Chlorella vulgaris and Scenedesmus quadricauda. J. Supercrit. Fluids 2016, 107, 408–413. [Google Scholar] [CrossRef]
  163. Peng, G.; Vogel, F.; Refardt, D.; Ludwig, C. Catalytic supercritical water gasification: Continuous methanization of Chlorella vulgaris. Ind. Eng. Chem. Res. 2017, 56, 6256–6265. [Google Scholar] [CrossRef]
  164. Hu, Z.; Ma, X.; Jiang, E. The effect of microwave pretreatment on chemical looping gasification of microalgae for syngas production. Energy Convers. Manag. 2017, 143, 513–521. [Google Scholar] [CrossRef]
  165. Liu, G.; Liao, Y.; Wu, Y.; Ma, X. Synthesis gas production from microalgae gasification in the presence of Fe2O3 oxygen carrier and CaO additive. Appl. Energy 2018, 212, 955–965. [Google Scholar] [CrossRef]
  166. Azizi, K.; Keshavarz Moraveji, M.; Abedini Najafabadi, H. A review on bio-fuel production from microalgal biomass by using pyrolysis method. Renew. Sustain. Energy Rev. 2018, 82, 3046–3059. [Google Scholar] [CrossRef]
  167. Wang, X.; Sheng, L.; Yang, X. Pyrolysis characteristics and pathways of protein, lipid and carbohydrate isolated from microalgae Nannochloropsis sp. Bioresour. Technol. 2017, 229, 119–125. [Google Scholar] [CrossRef] [PubMed]
  168. Lu, Q.; Guo, H.; Zhou, M.; Cui, M.; Dong, C.; Yang, Y. Selective preparation of monocyclic aromatic hydrocarbons from catalytic cracking of biomass fast pyrolysis vapors over Mo2N/HZSM-5 catalyst. Fuel Process. Technol. 2018, 173, 134–142. [Google Scholar] [CrossRef]
  169. Zainan, N.H.; Srivatsa, S.C.; Li, F.; Bhattacharya, S. Quality of bio-oil from catalytic pyrolysis of microalgae Chlorella vulgaris. Fuel 2018, 223, 12–19. [Google Scholar] [CrossRef]
  170. Zabed, H.M.; Qi, X.; Yun, J.; Zhang, H. Anaerobic digestion of microalgae biomass for methane production. In Microalgae Biotechnology for Development of Biofuel and Wastewater Treatment; Alam, M.A., Wang, Z., Eds.; Springer: Singapore, 2019; pp. 397–421. [Google Scholar]
  171. Wang, Y.; Guo, W.; Cheng, C.L.; Ho, S.H.; Chang, J.S.; Ren, N. Enhancing bio-butanol production from biomass of Chlorella vulgaris JSC-6 with sequential alkali pretreatment and acid hydrolysis. Bioresour. Technol. 2016, 200, 557–564. [Google Scholar] [CrossRef]
  172. Abomohra, A.E.F.; Elshobary, M. Biodiesel, bioethanol, and biobutanol production from microalgae. In Microalgae Biotechnology for Development of Biofuel and Wastewater Treatment; Alam, M.A., Wang, Z., Eds.; Springer: Singapore, 2019; pp. 293–321. [Google Scholar]
  173. Passos, F.; Carretero, J.; Ferrer, I. Comparing pretreatment methods for improving microalgae anaerobic digestion: Thermal, hydrothermal, microwave and ultrasound. Chem. Eng. J. 2015, 279, 667–672. [Google Scholar] [CrossRef]
  174. Bohutskyi, P.; Betenbaugh, M.J.; Bouwer, E.J. The effects of alternative pretreatment strategies on anaerobic digestion and methane production from different algal strains. Bioresour. Technol. 2014, 155, 366–372. [Google Scholar] [CrossRef]
  175. Passos, F.; Hom-Diaz, A.; Blanquez, P.; Vicent, T.; Ferrer, I. Improving biogas production from microalgae by enzymatic pretreatment. Bioresour. Technol. 2016, 199, 347–351. [Google Scholar] [CrossRef]
  176. Hernandez, D.; Riano, B.; Coca, M.; Garcia-Gonzalez, M.C. Saccharification of carbohydrates in microalgal biomass by physical, chemical and enzymatic pre-treatments as a previous step for bioethanol production. Chem. Eng. J. 2015, 262, 939–945. [Google Scholar] [CrossRef]
  177. Shokrkar, H.; Ebrahimi, S.; Zamani, M. Bioethanol production from acidic and enzymatic hydrolysates of mixed microalgae culture. Fuel 2017, 200, 380–386. [Google Scholar] [CrossRef]
  178. Cheng, L.M.; Lee, K.T.; Chan, D.J.C. Synergistic effect of pretreatment and fermentation process on carbohydrate-rich Scenedesmus dimorphus for bioethanol production. Energy Convers. Manag. 2017, 141, 410–419. [Google Scholar] [CrossRef]
  179. Dickinson, S.; Mientus, M.; Frey, D.; Amini-Hajibashi, A.; Ozturk, S.; Shaikh, F.; Sengupta, D.; El-Halwagi, M.M. A review of biodiesel production from microalgae. Clean Technol. Environ. Policy 2017, 19, 637–668. [Google Scholar] [CrossRef]
  180. Jeevan Kumar, S.P.; Vijay Kumar, G.; Dash, A.; Scholz, P.; Banerjee, R. Sustainable green solvents and techniques for lipid extraction from microalgae: A review. Algal Res. 2017, 21, 138–147. [Google Scholar] [CrossRef]
  181. Lai, Y.S.; De Francesco, F.; Aguinaga, A.; Parameswaran, P.; Rittmann, B.E. Improving lipid recovery from Scenedesmus wet biomass by surfactant-assisted disruption. Green Chem. 2016, 18, 1319–1326. [Google Scholar] [CrossRef] [Green Version]
  182. Orr, V.C.A.; Plechkova, N.V.; Seddon, K.R.; Rehmann, L. Disruption and wet extraction of the microalgae Chlorella vulgaris using room-temperature ionic liquids. ACS Sustain. Chem. Eng. 2016, 4, 591–600. [Google Scholar] [CrossRef]
  183. Lee, O.K.; Oh, Y.K.; Lee, E.Y. Bioethanol production from carbohydrate-enriched residual biomass obtained after lipid extraction of Chlorella sp. KR-1. Bioresour. Technol. 2015, 196, 22–27. [Google Scholar] [CrossRef]
  184. El-Dalatony, M.M.; Salama, E.S.; Kurade, M.B.; Kim, K.Y.; Govindwar, S.P.; Kim, J.R.; Kwon, E.E.; Min, B.; Jang, M.; Oh, S.E.; et al. Whole conversion of microalgal biomass into biofuels through successive high-throughput fermentation. Chem. Eng. J. 2019, 360, 797–805. [Google Scholar] [CrossRef]
  185. Chen, C.Y.; Liu, C.C. Optimization of lutein production with a two-stage mixotrophic cultivation system with Chlorella sorokiniana MB-1. Bioresour. Technol. 2018, 262, 74–79. [Google Scholar] [CrossRef]
  186. Xie, Y.; Zhao, X.; Chen, J.; Yang, X.; Ho, S.H.; Wang, B.; Chang, J.S.; Shen, Y. Enhancing cell growth and lutein productivity of Desmodesmus sp F51 by optimal utilization of inorganic carbon sources and ammonium salt. Bioresour. Technol. 2017, 244, 664–671. [Google Scholar] [CrossRef] [PubMed]
  187. European Commission. Communication from the Commission to the European Parliament, the Council, the European Economic and Social Committee and the Committee of the Regions. Closing the Loop—An EU Action Plan for the Circular Economy; European Commission: Brussels, Belgium, 2015. [Google Scholar]
  188. Bertolucci Paes, L.A.; Bezerra, B.S.; Deus, R.M.; Jugend, D.; Gomes Battistelle, R.A. Organic solid waste management in a circular economy perspective—A systematic review and SWOT analysis. J. Clean. Prod. 2019, 239, 118086. [Google Scholar] [CrossRef]
  189. Teigiserova, D.A.; Hamelin, L.; Thomsen, M. Review of high-value food waste and food residues biorefineries with focus on unavoidable wastes from processing. Resour. Conserv. Recycl. 2019, 149, 413–426. [Google Scholar] [CrossRef]
  190. Cao, Y.; Chen, S.S.; Zhang, S.C.; Ok, Y.S.; Matsagar, B.M.; Wu, K.C.W.; Tsang, D.C.W. Advances in lignin valorization towards bio-based chemicals and fuels: Lignin biorefinery. Bioresour. Technol. 2019, 291, 121878. [Google Scholar] [CrossRef] [PubMed]
  191. Mitra, M.; Mishra, S. Multiproduct biorefinery from Arthrospira spp. towards zero waste: Current status and future trends. Bioresour. Technol. 2019, 291, 121928. [Google Scholar] [CrossRef] [PubMed]
  192. Hemalatha, M.; Sarkar, O.; Mohan, S.V. Self-sustainable azolla-biorefinery platform for valorization of biobased products with circular-cascading design. Chem. Eng. J. 2019, 373, 1042–1053. [Google Scholar] [CrossRef]
  193. Milledge, J.J.; Nielsen, B.V.; Maneein, S.; Harvey, P.J. A brief review of anaerobic digestion of algae for bioenergy. Energies 2019, 12, 1166. [Google Scholar] [CrossRef] [Green Version]
  194. Abomohra, A.E.; Zheng, X.; Wang, Q.; Huang, J.; Ebaid, R. Enhancement of biodiesel yield and characteristics through in-situ solvo-thermal co-transesterification of wet microalgae with spent coffee grounds. Bioresour. Technol. 2021, 323, 124640. [Google Scholar] [CrossRef]
  195. Ma, X.; Zheng, H.; Addy, M.; Anderson, E.; Liu, Y.; Chen, P.; Ruan, R. Cultivation of Chlorella vulgaris in wastewater with waste glycerol: Strategies for improving nutrients removal and enhancing lipid production. Bioresour. Technol. 2016, 207, 252–261. [Google Scholar] [CrossRef]
  196. Abomohra, A.E.F.; Eladel, H.; El-Esawi, M.; Wang, S.; Wang, Q.; He, Z.X.; Feng, Y.Q.; Shang, H.; Hanelt, D. Effect of lipid-free microalgal biomass and waste glycerol on growth and lipid production of Scenedesmus obliquus: Innovative waste recycling for extraordinary lipid production. Bioresour. Technol. 2018, 249, 992–999. [Google Scholar] [CrossRef]
  197. Yarnold, J.; Karan, H.; Oey, M.; Hankamer, B. Microalgal Aquafeeds as Part of a Circular Bioeconomy. Trends Plant Sci. 2019, 24, 959–970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Chew, K.W.; Chia, S.R.; Yen, H.W.; Nomanbhay, S.; Ho, Y.C.; Show, P.L. Transformation of biomass waste into sustainable organic fertilizers. Sustainability 2019, 11, 2266. [Google Scholar] [CrossRef] [Green Version]
  199. Abad-Segura, E.; De La Fuente, A.B.; González-Zamar, M.; Belmonte-Ureña, L.J. Effects of circular economy policies on the environment and sustainable growth: Worldwide research. Sustainability 2020, 12, 5792. [Google Scholar] [CrossRef]
Figure 1. Electrochemical CO2R in heterogeneous (left) and homogeneous (right) catalyst systems.
Figure 1. Electrochemical CO2R in heterogeneous (left) and homogeneous (right) catalyst systems.
Sustainability 13 06962 g001
Figure 2. Current density of FA, CO, and H2 production in electrocatalytic CO2R over modified Cu electrode. Replotted data from Buckley et al. (2019) [35].
Figure 2. Current density of FA, CO, and H2 production in electrocatalytic CO2R over modified Cu electrode. Replotted data from Buckley et al. (2019) [35].
Sustainability 13 06962 g002
Figure 3. Faradaic efficiency of CO, FA, and H2 production in electrocatalytic CO2R over boron-doped diamond (BDD) cathodes as a function of operation potential. Replotted data from Tomisaki et al. (2019) [37].
Figure 3. Faradaic efficiency of CO, FA, and H2 production in electrocatalytic CO2R over boron-doped diamond (BDD) cathodes as a function of operation potential. Replotted data from Tomisaki et al. (2019) [37].
Sustainability 13 06962 g003
Figure 4. Illustration of CO2R pathways occurring at the protic and aprotic surface. Rearranged from Buckley et al. (2019) [35].
Figure 4. Illustration of CO2R pathways occurring at the protic and aprotic surface. Rearranged from Buckley et al. (2019) [35].
Sustainability 13 06962 g004
Figure 5. Biosequestration of CO2 by microalgae and further bio-refinery to valuable products.
Figure 5. Biosequestration of CO2 by microalgae and further bio-refinery to valuable products.
Sustainability 13 06962 g005
Figure 6. Conceptual presentation on the role of biotic CO2R in the circular bioeconomy.
Figure 6. Conceptual presentation on the role of biotic CO2R in the circular bioeconomy.
Sustainability 13 06962 g006
Table 1. Summary of CO2R performances included in this study.
Table 1. Summary of CO2R performances included in this study.
Base of CatalystFE (%) in FACO2R ConditionReference
Cu modified with polymeric38–45−0.7 VRHE in 0.05 M K2CO3 and 4 mM KCl with 5 sccm CO2[35]
boron-doped diamond~70−2.1 V (vs. Ag/AgCl) in KCl aqueous solution[37]
polycrystalline Sn~70−1.0 VRHE in 0.1 M KHCO3with 20 sccm CO2[43]
sulfur-modified copper~80−0.8 VRHE in 0.1 M KHCO3with 20 sccm CO2[44]
Cu~85−1.6 V vs. ferrocenium voltage in CO2-saturated [EMIM](BF4)/H2O (92/8 v/v) ionic liquid solution[45]
Cu (1.5 cm × 3 cm)~20−0.8 VRHE in 0.1 M KHCO3 with 20 sccm CO2[47]
Cu~20−1.4 VRHE in 0.5 M KCl with 70 sccm CO2[48]
Cu2O@Cu~40−0.7 VRHE in 0.1 M KHCO3 with 5 sccm CO2[51]
Iron-graphite electrode pair~18−0.6 VAg/AgCl with CO2 saturated 0.5 MNaHCO3 in anaerobic sludge digestion process[54]
Co incovalentorganic frameworks 1.02   mmol   h 1 g 1 in COMeCN with triethanolamineas sacrificial reducing agent and Ru(bpy)3Cl2·asphotosensitizerunder simulated sunlight[55]
In0.6Bi0.2Sn0.2 alloy on a halide perovskite~95−1.3 VRHE in 0.5 M KHCO3with 20 sccmCO2under simulated sunlight[59]
Cu0.5Zn0.5~60−1.3 VRHE in CO2 saturated 0.1 M KHCO3 under simulated sunlight[60]
SnOx/Sn~80−1.7 VAg/AgCl in 0.1 M CO2-saturated KHCO3[61]
Pd15Ni85~50−0.5 VRHE in 1.0 M CO2-saturated KHCO3[62]
Pb modified with 4-aminomethylbenzene~80−1.3 VRHE in 1.0 M CO2-saturated KHCO3[28]
N,S-codoped carbon catalysts~90 in CO−0.6 VRHE in 0.1 M KHCO3with 34 sccm CO2[65]
Pd85Cu15/C~86 in CO−0.9 VRHE in 1.0 M CO2-saturated KHCO3[66]
defective
β-Bi2O3 double-walled nanotubes
~90−0.8 VRHE in 0.5M KHCO3with 20 sccm CO2[67]
3.7 nm Pd nanoparticles~90 in CO−0.9 VRHE in 1.0 M CO2-saturated KHCO3[68]
boron-doped Pd catalyst~70−0.5 VRHE in 1.0 M CO2-saturated KHCO3[69]
Table 2. Performance of the selected catalysts for CO2 hydrogenation to CO and CH4.
Table 2. Performance of the selected catalysts for CO2 hydrogenation to CO and CH4.
CatalystT (°C)P (MPa)H2/CO2/InertGHSV
(mL g−1 h−1)
XCO2 (%)SCO (%)SCH4 (%)Reference
Ru/MnOx3000.122/5/73150,000251090[78]
Ru/Al2O332694
Ru/CeO283199
Ru/ZnO1946
PtCo/TiO23000.167/33/036,0008.2991[82]
PtCo/CeO29.1928
PtCo/ZrO27.89011
Co/ZrO24003.080/20/0360092.5< 1> 99[81]
Co/SiO280.1298
Co/Al2O377.8397
Co/TiO230.9964
Ni/ZIF-8 a4200.180/20/015,00043.8973[81]
Fe/ZIF-8 a43.8973
Ni/Fe/ZrO22300.580/20/05000821486[83]
γ-Fe2O34000.120/0.1/801,500,000453070[84]
Ni/MCM b4000.180/20/090,00073.2892[85]
a ZIF-8, a zinc-based zeoliticimidazolate framework, is pyrolyzed before metal doping; b MCM as zeolite.
Table 3. Performance of the selected catalysts for CO2 hydrogenation to methanol.
Table 3. Performance of the selected catalysts for CO2 hydrogenation to methanol.
CatalystT (°C)P (MPa)CO2/H2/InertGHSV
(mL g−1 h−1)
XCO2 (%)SCH3OH (%)Reference
Cu/SiO22504.172/24/436002.815[100]
Cu/ZnO1800.790/10/040000.994[97]
Cu/ZnO/ZrO2/SiO22402.030/90/1039,0005.238[99]
Cu/ZnO/ZrO2/MgO/Al2O32502.075/25/0200012.136[93]
Pd/ZnO2502.025/75/0360010.760[101]
Pd/SiO22504.172/24/436003.023[100]
Pd-Cu/SiO26.634
Pd-Cu/SiO22505.075/25/030,0001.627[102]
MnOx-Co3O42501.060/20/20120,00045.122[103]
In2O32704.060/20/2015,0001.155[76]
In2O33304.060/20/2015,0007.140
In2O3/ZrO23005.080/20/016,0005.2>99[104]
Table 4. Performance of the selected catalysts for CO2 hydrogenation to hydrocarbons.
Table 4. Performance of the selected catalysts for CO2 hydrogenation to hydrocarbons.
CatalystT (°C)P (MPa)H2/CO2/InertGHSV
(mL g−1 h−1)
XCO2 (%)SCO (%)Hydrocarbon Distribution (%)Reference
CH4C2=-C4=C2-C4C5+
Fe3O4320372/24/42000291760<1363[110]
Na-Fe3O4320372/24/4200041141647830
Fe2O33501.570/23/7115023211882[107]
K-Fe3O4/Fe2O3270573/25/2270037142442929[106]
CuFeO2300175/25/018001832431560[109]
CuFe2O4300175/25/0180016283814911
K-Fe-Co/Al2O33001.172/24/470031181627651[111]
Co/Fe oxide a2700.972/28/0200027148215<1[77]
Na-Co/Fe oxide a2700.972/28/02000234260292
Pyrolyzed
Fe-MIL-88B b
400375/25/03600461832231827[108]
K-pyrolyzed Fe-MIL-88B b400375/25/0360043263233619
In2O3/HZSM340373/24/39000134512079[112]
Ga2O3/HZSM340373/24/3900098653561
Fe2O3/HZSM340373/24/3900077422871
Na-Fe3O4/HZSM350372/24/44000332681874[105]
In2O3-ZrO2/SAPO400373/24/390003685476173[75]
a Pre-carburized under CO stream at 250 °C; b Fe-MIL-88B is an iron-based metal organic framework.
Table 5. Performance of the selected catalysts for CO2 hydrogenation to hydrocarbons.
Table 5. Performance of the selected catalysts for CO2 hydrogenation to hydrocarbons.
Microalgae SpeciesCO2 Conc. (v/v%)CO2 Fixation Rate
(g-CO2 L−1 d−1)
Biomass Production Rate
(g-DCW L−1 d−1)
Specific Growth Rate (d−1)Cultivation SystemCO2 Injection Rate (vvm), ModeLight Intensity (umol m−2 s−1), Light/Dark CycleReference
Acutodesmusobliquus14.10.22 b0.121.09Flask1.66, intermittent120, 24/0[124]
Aphanothecemicroscopica Nägeli15 a14.14 (from carbon balance)0.471.4Bubble column1, intermittent150, 24/0[140]
Asterarcysquadricellulare5 a0.39 b0.211.2Flask0.5, continuous250, 14/10[123]
Chlorella minutissima3.51.170.641.66Airlift0.43, continuous50~260 f, 24/0[139]
Chlorella sorokiniana5 a0.39 b0.222.42Flask0.5, continuous250, 14/10[123]
Chlorella sorokiniana15.60.310.170.46Airlift0.33, continuous120, 24/0[121]
Chlorella sp.11.000.530.615Flat-panel0.25, continuous130, 24/0[131]
Chlorella sp.20.77 b0.420.62Bubble column0.2, continuous400, 24/0[115]
Chlorella sp.30.460.31NABubble column-, continuous115, 24/0[119]
Chlorella sp.51.650.870.15Bubble column0.5, continuous56, 24/0[133]
Chlorella sp.50.100.060.25FlaskNo aeration450, 24/0[141]
Chlorella sp.8 a2.330.841.11Bubble column0.2, continuous300, 24/0[136]
Chlorella sp. c101.330.730.55Bubble column0.2, intermittent300, 24/0[126]
Chlorella sp.12.5 a0.97 b0.530.827Bubble column0.2, continuous300, 24/0[142]
Chlorella sp. d300.61 b0.33NABubble column0.67, continuous95, 24/0[128]
Chlorella vulgaris2.53.712.06NABubble column0.36, continuous300, 18/6[138]
Chlorella vulgaris e30.310.17NAFlaskNo aeration40, 12/12[88]
Chlorella vulgaris51.32 b0.72NABubble column0.1, continuous160, 24/0[143]
Chlorella vulgaris50.100.121.36Flask0.5, continuous160, 24/0[118]
Chlorella vulgaris100.430.260.46Tubular0.5, continuous100, 12/12[120]
Chlorella vulgaris100.31 b0.171.34 cFlask-, intermittent40, 14/10[134]
Desmodesmus sp.2.51.580.88NATubular0.2, continuous300~1000 f, 24/0[117]
Microcystisaeruginosa g50.110.141.5Flask0.5, continuous160, 24/0[118]
Pseudokirchneriellasubcapitata50.090.100.89Flask0.5, continuous160, 24/0[118]
Scenedesmusdimorphus15 a0.890.49NATubular0.25, intermittent100, 24/0[132]
Scenedesmus sp.150.17 b0.09NAFlaskNo aeration-, 16/8[122]
Synechocystissalina h50.110.141.65Flask0.5, continuous160, 24/0[118]
Tetraselmissuecica,50.110.070.29FlaskNo aeration450, 24/0[141]
Chlorella sp., Scenedesmus sp., Sphaerocystis sp., Spirulina sp.200.27 b0.15NAFlaskNo aeration80, 16/8[125]
a Synthetic or real flue gas; b Estimated by assuming that 1.83 g of CO2 is fixed in every gram of microalgal biomass (Cheah et al., 2015) [112]; c Chemically mutated strain by N-methyl-N′-nitro-N-nitrosoguanidine; d Evolved under 10% CO2 after 31 adaptive laboratory evolution cycles; e Genetically modified strain; f Between 0~36 h; g Cyanobacteria; h Increased linearly.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, T.; Dong, C.-D.; Lin, J.-Y.; Chen, C.-W.; Chang, J.-S.; Kim, H.; Huang, C.-P.; Hung, C.-M. Recent Advances in Carbon Dioxide Conversion: A Circular Bioeconomy Perspective. Sustainability 2021, 13, 6962. https://0-doi-org.brum.beds.ac.uk/10.3390/su13126962

AMA Style

Wang T, Dong C-D, Lin J-Y, Chen C-W, Chang J-S, Kim H, Huang C-P, Hung C-M. Recent Advances in Carbon Dioxide Conversion: A Circular Bioeconomy Perspective. Sustainability. 2021; 13(12):6962. https://0-doi-org.brum.beds.ac.uk/10.3390/su13126962

Chicago/Turabian Style

Wang, TsingHai, Cheng-Di Dong, Jui-Yen Lin, Chiu-Wen Chen, Jo-Shu Chang, Hyunook Kim, Chin-Pao Huang, and Chang-Mao Hung. 2021. "Recent Advances in Carbon Dioxide Conversion: A Circular Bioeconomy Perspective" Sustainability 13, no. 12: 6962. https://0-doi-org.brum.beds.ac.uk/10.3390/su13126962

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