Next Article in Journal
Comprehensive Study on Environmental Behaviour and Degradation by Photolytic/Photocatalytic Oxidation Processes of Pharmaceutical Memantine
Next Article in Special Issue
Applications of Brewer’s Spent Grain Hemicelluloses in Biorefineries: Extraction and Value-Added Product Obtention
Previous Article in Journal
Evidence of Synergy Effects between Zinc and Copper Oxides with Acidic Sites on Natural Zeolite during Photocatalytic Oxidation of Ethylene Using Operando DRIFTS Studies
Previous Article in Special Issue
Design of a New Chemoenzymatic Process for Producing Epoxidized Monoalkyl Esters from Used Soybean Cooking Oil and Fusel Oil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on the Progress in Chemo-Enzymatic Processes for CO2 Conversion and Upcycling

1
Department of Chemical & Petroleum Engineering, UCSI University, Kuala Lumpur 56000, Selangor, Malaysia
2
Faculty of Medicine & Health Sciences, UCSI University, Kuala Lumpur 56000, Selangor, Malaysia
3
NUS Environmental Research Institute, National University of Singapore, 1 Create Way, Singapore 138602, Singapore
4
Department of Biological Sciences, School of Medical and Life Sciences, Sunway University, No. 5, Jalan University, Bandar Sunway, Petaling Jaya 47500, Selangor, Malaysia
5
Graphene and Advanced 2D Materials Research Group (GAMRG), School of Engineering and Technology, Sunway University, No. 5, Jalan University, Bandar Sunway, Petaling Jaya 47500, Selangor, Malaysia
6
School of Applied and Life Sciences, Uttaranchal University, Dehradun 248007, Uttarakhand, India
7
Amity Institute of Biotechnology, Amity University Jharkhand, Ranchi 201313, Jharkhand, India
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(3), 611; https://doi.org/10.3390/catal13030611
Submission received: 31 January 2023 / Revised: 12 March 2023 / Accepted: 13 March 2023 / Published: 17 March 2023
(This article belongs to the Special Issue New Advances in Chemoenzymatic Synthesis)

Abstract

:
The increasing concentration of atmospheric CO2 due to human activities has resulted in serious environmental issues such as global warming and calls for efficient ways to reduce CO2 from the environment. The conversion of CO2 into value-added compounds such as methane, formic acid, and methanol has emerged as a promising strategy for CO2 utilization. Among the different techniques, the enzymatic approach based on the CO2 metabolic process in cells presents a powerful and eco-friendly method for effective CO2 conversion and upcycling. This review discusses the catalytic conversion of CO2 using single and multienzyme systems, followed by various chemo-enzymatic processes to produce bicarbonates, bulk chemicals, synthetic organic fuel and synthetic polymer. We also highlight the challenges and prospects for future progress in CO2 conversion via chemo-enzymatic processes for a sustainable solution to reduce the global carbon footprint.

1. Introduction

Economic development activities such as industrialization and urbanization have resulted in significant CO2 production which remains a major factor for global warming, ocean acidification and a dreadful pollutant affecting human health. The predominant source of anthropogenic CO2 emissions is the combustion of fossil fuels for energy generation in residential, industrial and transportation sectors. For example, burning 1 ton of carbon in fossil fuels produces over 3.5 tons of CO2 [1]. Atmospheric CO2 concentration was reported to be 405 parts per million (ppm) in 2017 and is projected to increase to 570 ppm by the end of the century [2]. In light of these circumstances, the International Energy Agency (IEA) has recommended restricting annual CO2 emissions to <15 gigatonnes to limit the rise in temperature to within 2 °C by 2050 [3,4].
Despite the rising concerns about CO2 emission, it remains a challenge to propose a truly low-carbon technology to the public market within a short time frame to reduce CO2 emission to a safer level in the atmosphere [5]. As such, the best approach to address the issue will be capturing and converting CO2 into useful products or compounds; i.e., CO2 utilization. Today, various techniques involving chemical, photochemical, electrochemical and enzymatic methods have been reported in the literature to catalyze CO2 fixation/conversion reactions to capture and store the increasing concentration of CO2 from the environment as an efficient energy source, or to produce value-added hydrocarbons such as methane, formic acid and methanol [6,7,8,9]. Among the aforementioned techniques, chemical, photochemical and electrochemical methods are limited by their low selectivity due to the abundant and stable form of carbon element in CO2 molecules and the challenges to attain a high-performing catalyst [10,11,12].
On the other hand, an enzymatic method is an eco-friendly approach which has been reported to be efficient for CO2 fixation/conversion reactions owing to their enhanced stereo-specificity and region-chemo selectivity. For instance, the enzymatic technique using three dehydrogenases was first reported by Obert and Dave [13], where the overall process involves the initial reduction of CO2 to formate catalyzed by formate dehydrogenase (FDH); followed by the reduction of formate to formaldehyde by formaldehyde dehydrogenase (FaDH); and finally, the reduction of formaldehyde to methanol by alcohol dehydrogenase (ADH). In the entire process, the reduced nicotinamide adenine dinucleotide (NADH) acts as the terminal donor for all the dehydrogenase-catalyzed reductions. Since methanol emits less CO2 after combustion and has high volumetric and gravimetric energy density, it exhibits significant potential to replace fossil fuels in the future. The enzymatic route for the CO2 to methanol conversion reaction has been reported to be selective and to exhibit higher reaction rates under mild conditions [14,15]. Other studies have reported that issues concerning the stability of the enzyme can be addressed by immobilizing the enzyme in ultrathin, hybrid microcapsules of catechol and gelatin [16,17].
Whilst much of the emphasis has been intensifying on experimental studies of an enzymatic method for efficient CO2 conversion reactions, comprehensive reviews concerning reaction routes and mechanisms for catalytic conversion of CO2 by single and multienzyme systems remain limited. Herein, the authors attempt to gather information from the earliest to the most recent developments regarding the catalytic conversion of CO2 by single and multienzyme systems; and the subsequent upcycling of CO2 via various chemo-enzymatic processes such as the production of bicarbonates, bulk chemicals, synthetic organic fuel and synthetic polymer. Lastly, we highlight the challenges and prospects for future progress in CO2 conversion reactions from the chemo-enzymatic processes.

2. Overview of Chemical Methods for CO2 Conversion

2.1. Electrochemical

The electrochemical method involves using a cathode and an anode, which supply electricity to the electrolyte in a setup illustrated in Figure 1A. An ion exchange membrane separates the electrodes, and the electrode consists of catalyst and substrate layers to facilitate the transfer of CO2 reactant gas and the product. As a result, it is essential to ensure effective mass transfer between the electrodes. In a separate investigation conducted by Yang et al., a novel three-compartment electrochemical cell setup (Figure 1B) was developed, allowing pure formic acid to produce within a concentration range of 5–20 wt% [18]. The improved and consistent electrochemical performance for up to 500 h was attributed to using an electrochemical cell incorporating a Dioxide Materials Sustainion™ anion exchange membrane and a nanoparticle Sn GDE cathode containing an imidazole ionomer.
However, Hori et al. [19] showed the drawbacks of planar electrodes due to mass transfer limitation due to the poor solubility of CO2. It has been reported that using gas diffusion electrodes as a cathode in combination with an improved catalyst deposition layer led to improvements in the electrode performance [20]. Due to their distinctive electrochemical properties, many studies have been carried out by utilizing copper sulfide compounds as a catalyst in the CO2 electrochemical reduction process [21,22]. However, it should be noted that the electrochemical approach requires a high cost of electricity for the reaction.
In this context, current research on lowering overpotentials has been carried out to lower the high electricity cost. For instance, Rosen et al. [23] reported CO2 could be reduced to carbon monoxide (CO) at the applied voltage of 1.5 V to a very low overpotential of <0.2 V under Pt/Nafion/(Ag, EMIM-BF4) electrocatalytic system. The system continuously produced CO for >7 h at Faradaic efficiencies >96%. Another example was the electrocatalytic reduction of CO2 by carbon monoxide dehydrogenase (CODH), a naturally occurring enzyme [24]. Shin et al. [25] achieved a current efficiency of 100% at −0.57 V vs. NHE in a 0.1 M phosphate buffer (pH 6.3) with no overpotential with CODH catalysts. Apart from CO production, methanol can be produced. The electrocatalyst containing histamine generates formic acid, an intermediary compound that subsequently reacts with Pt nanoparticles. This combination enhances the availability of adsorbed hydrogen atoms, producing highly hydrogenated products, such as methanol [26]. Further, the selectivity of the desired product can be increased by using the appropriate electrolyte, catalyst and potential applied to the process.
Figure 1. (A) Basic configuration of electrochemical reduction of CO2 conversion consisting of cathode and anode through which electricity is applied in the electrolytic solution [27]. Reproduced with permission from Elsevier 2021. (B) Electrochemical conversion of CO2 to formic acid facilitated by the presence of imidazolium ionomer in the nanoparticle Sn catalyst at GDE cathode [18]. Reproduced with permission from Elsevier 2017.
Figure 1. (A) Basic configuration of electrochemical reduction of CO2 conversion consisting of cathode and anode through which electricity is applied in the electrolytic solution [27]. Reproduced with permission from Elsevier 2021. (B) Electrochemical conversion of CO2 to formic acid facilitated by the presence of imidazolium ionomer in the nanoparticle Sn catalyst at GDE cathode [18]. Reproduced with permission from Elsevier 2017.
Catalysts 13 00611 g001

2.2. Thermal

Various catalytic metal particles have been reported for CO2 conversion. Incorporating silica or alumina support has also been attributed to increased catalytic activity-selectivity performance. Specifically, Cu particles (at variable loadings and particle sizes, including the nanoscales) supported with ZnO have been extensively explored for methanol production. For instance, Huang et al. [28] modified a range of CuO/ZnO/ZrO2 with ZrO2 support and tested at 240 °C and 3 MPa pressure. Results revealed both CO2 conversion and CH3OH selectivity were appreciably enhanced by the ZrO2 incorporation. Din et al. [29] utilized nano-structured materials developed by supporting the Cu particles with ZrO2. Liang et al. [30] utilized another porous and structured system based on Cu–Zn/Al-foams having monolithic configurations. The catalytic choice with monolithic systems in micro-reactors was shown to be promising in enhancing catalyst stability by hindering coking and elevating reaction activity and thus, the production yield. Other groups of catalysts, such as Pd and Fe, were the alternatives. These findings could be correlated to the works reported by Collins et al. [31] and Bonivardi et al. [32], who employed supported Pd–Ga materials under variable conditions. Their results highlighted that the modification of the catalyst enhanced H2 activation. Another group of catalysts with proficiency is the Fe-based systems. Such work was published by Dorner et al. [33], where CeO2-modified Fe/Mn/Alumina catalysts were developed. Results revealed that CO2 conversion of >14% and CH3OH selectivity >70% could be realized with enhanced stability based on loadings <10 wt%. Over 10 wt% showed the activity/selectivity decayed substantially. Numerous studies have also been reported with other metals, such as La, Mg, Zn, etc., associated with supported Fe particles. Unfortunately, the results were not satisfying (low quality), probably due to poor Fe active site interactions that hindered CO2/H2 adsorption [34]. However, in certain cases, positive effects of sintering inhibition were achieved depending on the support material (i.e., in the order Al2O3 > TiO2 > SiO2) [35,36]. Metal carbides have also been highlighted for CO2 conversion. For instance, Rodriguez et al. [37] showed that TiC could create effective support in situ, especially when loaded with active sites (Au, Ni or Cu). Asara et al. [38] corroborated that Au/TiC materials could yield good reaction properties when the TiC was well structured with improved active surface sites coupled to Au dispersion.

3. In Vitro Conversion of CO2

3.1. CO2 Conversion via Single Enzyme In Vitro

Enzyme-based CO2 conversion has gained significant importance due to its ability to perform complicated reactions and produces a high yield of desired fuels or chemicals. Table 1 shows the enzymes used to produce various products via in vitro CO2 conversion. CO2 conversion in vitro is also an important method for CO2 capture, sequestration and utilization, which can be accomplished by directly adopting a single enzyme as listed below:
  • Conversion of CO2 by oxidoreductases;
  • Conversion of CO2 by lyases;
  • Conversion of CO2 by synthases.
Among the three single enzyme in vitro reactions aforementioned, reactions performed by synthases–such as acetyl-CoA synthase, pyruvate synthase or acetyl-CoA carboxylase, among others–produce products that do not possess any significant importance in our daily life. As such, synthases can be considered unsuitable for in vitro applications. In comparison to other major routes, only oxidoreductases (i.e., formate dehydrogenase (FDH), Carbon dioxide reductase, Carbon monoxide dehydrogenase (CODH), Remodeled nitrogenase) and lyases (i.e., carbonic anhydrase, pyruvate decarboxylase, etc.) have been discovered from certain organisms which assisted in the synthesis of several fuels or chemicals such as formate, CO, methane or bicarbonate, among others. The following section will describe the CO2 conversion routes via a single enzyme in vitro; i.e., the conversion of CO2 by oxidoreductases and the conversion of CO2 by lyases.

3.1.1. Conversion of CO2 by Oxidoreductases

Generally, oxidoreductases are enzymes that catalyze the transfer of electrons from one reducer or electron donor molecule to another oxidant or electron acceptor. NADPH/NADP+ or NADH/NAD+ are usually used as cofactors [39,40]. The conversion of CO2 by oxidoreductases produces carbon-based energy sources by reducing the oxidation state of carbon element. Various products can be obtained from the conversion of CO2 catalyzed by oxidoreductases. For example, CO2 reduction produces formate (an important chemical and fuel) by formate dehydrogenase (FDH), as shown in Equation (1).
C O 2 + H + + 2 e H C O O
In the forward reaction, FDH catalyzes the reduction of CO2 to formate, while in the reverse reaction, it catalyzes the oxidation of formate to CO2. A reverse reaction is more likely to occur in the presence of a coenzyme (NADH) regeneration system catalyzed by FDH. Although most FDH prefer oxidation reactions, some FDH can reduce CO2 to formate that can be easily converted to other important chemicals without requiring other organic compounds. FDH can be classified as metal-independent; i.e., metal-free FDH enzymes catalyze reactions from CHOOH to CO2 using nicotinamide adenine dinucleotide (NAD+) as a cofactor, or metal-dependent enzymes based on molybdenum (Mo) or tungsten (W) metals, to reversibly reduce CO2 to CHOOH by applying a series of enzymes adsorbed from S. fumaroxidans on electrode surface which showed high efficiency in the catalytic reduction of CO2 to formate (Figure 2A) [41]. The rate of the reaction was highly sensitive to pH, where a six-fold increase was reported from pH 7.5 to 5.5 (Figure 2B).
However, ion-type cofactors such as NADH are costly and not easily available, restricting their application in the catalytic reduction of CO2 to formate. In view of these limitations, Beller et al. discovered bacterial hydrogen-dependent CO2 reductase (HDCR) from the acetogenic bacterium Acetobacterium woodie for the conversion of CO2 to formate [42,43]. The bacterium grows with H2 and CO2 at ambient conditions using ancient energy conservation pathways. The HDCR enzyme is composed of four components; i.e., putative formate dehydrogenase (FdhF1/2) and iron-iron hydrogenase (HydA2) as the major subunits surrounded by two small electron transfer subunits (HycB2/3), where FdhF1/2 acts as the catalyst to reduce CO2 to formate. At the same time, HydA2 activates/oxidizes H2 into two H+ accompanied by acquiring two electrons, while HycB2/3 is responsible for transferring the electrons from HydA2 to FdhF1/2. When utilized for the catalytic conversion of CO2 and H2 to formate, a very high reaction rate was reported; i.e., 10 mmol min−1 mg−1. Compared with other chemical catalysts, the hydrogenation of CO2 was promoted with an initial turnover frequency (TOF) of 70 h−1 [44]. On the other hand, the purified HDCR catalyzed a similar reaction at 30 °C with a TOF of 101 600 h−1, which was approximately 1500 times higher than the chemical catalysis.
Carbon monoxide (CO) is another important product that can be produced by reducing CO2 using carbon monoxide dehydrogenase (CODH). CO exhibits significant fuel value, which can be readily converted to methanol as liquid fuel and is an essential feedstock in various synthetic processes such as Fisher-Tropsch, Monsanto and Cativa processes as shown in Equation (2).
C O + H 2 O C O 2 + 2 H + + 2 e
The reduction of CO2 to CO can be achieved with the aid of two types of CO dehydrogenase catalyzing the reaction. Firstly, CO dehydrogenases containing Fe4S4Ni in the active center of oxygen-sensitive obligatory anaerobes such as Moorella thermoacetica, which was first reported by Shin and co-workers [25]. A significant advancement in their study was the combination of electrocatalysis/photocatalysis with the enzymatic conversion of CO2 to CO. These enzymes can oxidize CO at a rate of 40,000/s using methyl viologen cation as oxidant at 70 °C, and can reduce CO2 at a rate of 45/s while also exhibiting close thermodynamic potential for the CO2/CO pair. However, the electrocatalysis/photocatalysis conversion of CO2 requires significant overpotential, which consumes a lot of energy and might generate a mixture of products such as CO, methane and methanol. In order to address this concern, Armstrong et al. conducted multiple studies wherein immobilized NiFe CODH was evaluated for enzyme activity while rotating at high speed in an anaerobic sealed cell, using pyrolytic graphite as a substrate [45,46]. Authors reported that the standard reduction potential for interconversion between CO2 and CO was much higher (E = −0.46 V) than without a catalyst (E = −1.9 V), which ascertains the lower overpotential requirements while achieving high selectivity for yielding 100% CO. The second class of CODH contains air-stable MoSCu, commonly found in aerobes such as Oligotropha carboxidovorans, in the active center. Some studies have reported that glutamate (Glu 763) and phenylamine (Phe 390) are important for binding ligaments of Mo and Cu to perform their activity. In contrast, other studies have reported that these enzymes achieve a lower catalytic conversion rate; i.e., 100/s for CO oxidation [47,48].
Methane is another product that can be produced by converting CO2 using remodeled nitrogenase as described in Equations (3) and (4). In earlier studies by Yang et al., it was reported that nitrogenase was not capable of triggering the reaction to catalyze eight electron reduction of CO2 to methane since no methane was detected even after 20 min of reaction [49]. However, when several amino acids (i.e., α-195, α-70 or α-191) that approach FeMo-cofactor active sites (Figure 2C) that control substrate binding and reduction reaction were substituted in appropriate molecules, the catalytic behavior was altered (Figure 2D). The remodeling of nitrogenase by replacing α-195 with Gln and α-70 with Ala successfully catalyzed the formation of methane from CO2 where 16 nmol (methane) and nmol−1 (MoFe protein) were obtained in the 20 min of reaction.
N 2 + 8 H + + 16 A T P + 8 e 2 N H 3 + H 2 + 16 A D P + 16 P i
C O 2 + 8 H + + 8 e C H 4 + 2 H 2 O
Table 1. Enzymes used for production of various products via in vitro CO2 conversion (modified from [48]).
Table 1. Enzymes used for production of various products via in vitro CO2 conversion (modified from [48]).
ProductEnzymeAdvantagesLimitationsRef.
Bicarbonate (HCO3)Carbonic anhydraseRapid cycle
  • Unstable, cannot be reused.
  • Cannot obtain value-added products.
  • Poor activity in harsh conditions.
[50,51,52]
Formate (HCOO)Formate dehydrogenaseObtaining value-added products
  • Requires modification for electrocatalytic/photocatalytic application.
  • Favors forward reaction.
  • Expensive cofactor.
  • Requires improvement for long-term covalent immobilization.
[53,54,55]
Carbon dioxide reductase
  • Doesn’t require cofactor
  • Reaction direction easily controlled by substrate concentration
  • CO can easily inhibit reactions in cascade system
[42,43,56]
Methanol (CH3OH)Multiple dehydrogenases (FDH + FaDH + ADH)
  • Value added products.
  • Can be processed at moderate conditions
  • Not stable for long-term activity
  • Requires immobilization for industrial application.
[13,57,58]
Methane (CH4)Remodeled NitrogenaseNot inhibited by H2
  • O2 sensitive enzymes
  • Requires modification to improve turnover number and frequencies
[59,60]
Carbon Monoxide (CO)Carbon monoxide dehydrogenase
  • Air stable,
  • Product reaction not easily convertible to value-added products
  • Low stability
[45]
Figure 2. (A) Schematic of catalytic interconversion between CO2 and formate. (B) Kinetics of catalytic reaction by tungsten containing FDH as a function of pH value [41]. Copyright (2012) Yang et al. (C) FeMo cofactor with some key amino acid residues [49]. (D) CH4 formation as a function of time for different MoFe proteins [49]. Copyright (2008) National Academy of Sciences, USA.
Figure 2. (A) Schematic of catalytic interconversion between CO2 and formate. (B) Kinetics of catalytic reaction by tungsten containing FDH as a function of pH value [41]. Copyright (2012) Yang et al. (C) FeMo cofactor with some key amino acid residues [49]. (D) CH4 formation as a function of time for different MoFe proteins [49]. Copyright (2008) National Academy of Sciences, USA.
Catalysts 13 00611 g002

3.1.2. Conversion of CO2 by Lyases

Lyase is an enzyme that facilitates the cleavage of various chemical bonds using techniques that do not involve hydrolysis or oxidation but rather substitution reactions. This enzymatic activity often results in the creation of a new double bond or ring structure. Conversely, the reverse reaction, known as the Michael reaction, allows for the carboxylation of unprocessed molecules with carbon dioxide to produce valuable products. For example, carbonic anhydrase (CA) can convert CO2 to bicarbonate, as shown in Equation (5).
C O 2 + H 2 O H + + H C O 3
CA commonly exist in mammals, plants, algae or bacteria and plays the primary role in the interconversion between CO2 and bicarbonate to maintain acid-base balance in the blood and other tissues. Based on the origin of enzymes and their structural fold, several classes of CA are available, such as α, β γ, δ and ζ, all of which are composed of a tetrahedral active center with a metal ion such as divalent zinc or related metal ion [61,62]. For the reaction mechanism of CA, a hydroxide atom in the active site attacks a CO2 molecule and then reacts with the water molecule from the reaction environment to release protons (i.e., rate limiting step) [63]. Higher CO2 concentration directs to bicarbonate production. To date, CA catalytic conversion has been used for absorption, membrane separation, mineralization/precipitation for CO2 capture, sequestration and utilization. For example, the CA catalytic conversion of CO2 by chemical absorption using regenerable alkaline solvents increases the CO2 absorption rate in the alkaline aqueous solvents with lower heat and energy input for desorption [64]. Despite the advantages, the harsh conditions involved in the absorption process, such as high temperature (50–125 °C), high concentration of organic amines and trace contaminants, causes enzyme denaturation, which leads to reduced enzyme activity and stability [27].
Several modifications have been developed to overcome the aforementioned limitations, such as sourcing CA from thermophilic organisms, generating themo-resistant CA via protein engineering techniques, immobilizing CA on suitable supports or modifying the absorption process. For instance, in a study by Zhang et al., CA was immobilized on nonporous silica-based nanoparticles in a potassium carbonate-based absorption process to improve enzyme stability and chemical resistance [65]. Authors reported that immobilized CA retained 56–88% of its original activity for a 60-day period at 50 °C, in comparison to free CA, which retained only 30% of its original activity. Another study by Hwang et al. revealed using a CA-assisted method to create crystalline CaCO3 composites together with in situ CA entrapment [66]. This study largely used free CA as a catalyst to speed up the precipitation and mineralization processes. Authors reported that CA immobilized in CaCO3 composites retained 43% catalytic activity. Additional investigations have noted other immobilization carriers since CaCO3 is pH and temperature sensitive, which may prevent the application of CA in particular regions. For example, Forsyth et al. reported the green approach for the recyclable conversion of CO2 to CaCO3 catalyzed by CA encapsulated bio-inspired silica [67]. Authors highlighted that immobilized CA exhibited little enzyme leaching with excellent recycling, storage and thermal stabilities.

3.2. CO2 Conversion via Multienzyme In Vitro

In living cells, enzymes catalyze a wide variety of metabolic processes that involve multiple reaction steps. The formation of macromolecular enzyme complexes efficiently transfers the intermediates from one catalytic site to the other [68]. A multienzyme-based approach for CO2 conversion utilizing naturally occurring reactions may prove efficient and cost-effective. To date, at least six CO2 conversion pathways have been validated: the Calvin cycle, reductive tricarboxylic acid (TCA) cycle, reductive acetyl-CoA pathway, 3-hydroxypropionate bicycle, dicarboxylate/4-hydroxybutyrate cycle, and reductive glycine pathway; in a variety of living cells [69,70]. As a result, this process, known as metabolic channeling, has motivated researchers to artificially combine multienzyme biocatalysts to achieve the desired goal of creating bio-based compounds. Among them, using cell-free enzymes or immobilized multiple enzymes cascade systems for CO2 fixation is an up-and-coming area. These types of cascades have several benefits over cellular conversion. In such multienzyme cascade systems, the enzyme rate of reaction can be enhanced, and a large concentration of CO2 can be processed while recycling enzymes for several cycles. Reaction rates are efficiently optimized without using toxic substrates or reaction intermediates, thus facilitating metabolic pathways by regulating the activities of competing pathways [71]. Pyruvate dehydrogenase (PDH) complex, tryptophan synthase channel, fatty acid oxidation pathway and citric acid cycle, are metabolic channeling examples that mimic natural enzyme complexes artificially [72]. For instance, direct product transfer as a substrate from one enzyme to another nearby enzyme with the bulk phase without equilibration [71], as illustrated in Figure 3. Such a scenario happens when the distance between two active sites is short, and thus the overall reaction rate can be enhanced efficiently [73]. In this regard, different active sites can be closer spatially and assembled as a complex to provide efficient reaction control with multifunctional enzymes (multienzyme cascade systems). Moreover, these systems offer ease in separating and purifying the end product. This section reviews various CO2 conversion via multienzymes in vitro strategies which have been exploited so far.

3.2.1. Conversion of CO2 to Methanol by Multiple Dehydrogenases

One of the important products of CO2 conversion is methanol, for the following reasons: (i) the recycling of “greenhouse” gas, and (ii) the efficient production of sustainable/renewable fuel alternatives. In comparison to the fuels (CO, methane, etc.) produced from the single-enzyme route, the liquid methanol produced from multienzyme reactions has much higher energy capacity and is easier to transport. In early 2007, Bilal El-Zahab et al. demonstrated methanol production from CO2 by applying a group of enzymes, i.e., FDH, FaDH, and ADH. The reaction catalyzed by these enzymes was an NADH dependent reaction [74]. To carry out the in vitro reaction, they immobilized all the enzymes and the NADH covalently on the polystyrene microparticle (~500 nm) (Figure 4A). The reaction was conducted by bubbling the CO2 gas in the reaction mixture; and a promising improvement in cofactor utilization was observed during the production of methanol from CO2. On the other hand, another approach for methanol production was investigated by constructing an immobilized multienzyme cascade. For instance, the support matrix for the enzyme entrapment was fabricated as a hybrid microcapsule from catechol modified gelatin (GelC) [57]. Enzyme-GelC cascade was synthesized by systematic stepwise entrapment of three enzymes, FAD, FaDH, and alcohol dehydrogenase, via physical and covalent attachment by forming silica nanoparticle layer facilitating enzyme-GelC-Si cascade (Figure 4B). This microreactor showed an enhanced methanol yield of up to 71%, along with a high recyclability rate, activity, and stability strength. Similarly, another study was investigated by carrying out the co-immobilization of FDH, FaDH, and alcohol dehydrogenase (ADH) on the commercially available ultrafiltration (UF) flat-sheet polymeric membrane by employing a simple pressure filtration method [17]. Two distinct approaches were used; i.e., co-immobilization and sequential immobilization. For co-immobilization, a simple method of UF membrane fouling was used under pressure until the membrane was saturated with enzymes (Figure 4C). In sequential immobilization, three different UF membranes were saturated with enzymes sequentially and stacked onto each other.
In the past, enzymatic reductions for CO2 were limited because a sacrificial coenzyme, such as nicotinamide adenine dinucleotide (NADH), was required to provide electrons and the hydrogen equivalent. However, a new bio-electrocatalytic reduction method for converting CO2 to methanol was studied, which involves injecting electrons directly from electrodes into immobilized enzymes, thus eliminating the need for the coenzyme NADH [75]. To achieve this, a carbon-felt electrode was created by depositing FDH, FaDH, and ADH within an alginate-silicate hybrid gel (Figure 4D). Using this technique, the researchers were able to produce almost 0.15 ppm of methanol. This method of immobilizing enzymes on electrodes presents an opportunity for electrochemical applications in various reactions where an alternative to the costly sacrificial coenzyme NADH is desired [76].
A magnetic nanoparticle multienzyme cascade was developed by immobilizing ADH, FaDH, and FDH powered by NAD+/NADH and glutamate dehydrogenase (GDH) as the coenzyme regenerating system [77] (Figure 4E). The stepwise reaction scheme led to only a 2.3% yield of methanol per NADH. However, in a batch system under CO2 pressure, combining the four immobilized enzymes increased the methanol yield by 64-fold. Therefore, this study indicated a successful regeneration of NADH in situ, envisaging a real possibility of using immobilized enzymes to perform the cascade CO2-methanol reaction.

3.2.2. Conversion of CO2 to Other Fuels/Chemicals/Materials by Use of a Multienzyme System

Developing novel reaction routes for the direct production of advanced fuels/chemicals/materials (with higher number of carbon) would also be highly competitive. Recently, a relatively simpler multienzyme route for converting CO2 and ethanol into L-lactic acid denoted a novel and sustainable alternative to synthesizing building blocks for biodegradable polymers [78] (Figure 5A). The catalytic process is composed of three single-enzyme routes: (1) the oxidation of ethanol to acetaldehyde by ADH in the presence of NAD+, (2) the synthesis of pyruvate from CO2 and acetaldehyde catalyzed by pyruvate decarboxylase, and (3) the reduction of pyruvate to L-lactic acid by lactate dehydrogenase (LDH) in the presence of NADH. It is worth noting that the cycling of NAD+/NADH could be achieved via the first and third single-enzyme routes. Another encouraging advance in technology was reported on the photosynthetic recycling of CO2 into isobutyraldehyde [79] (Figure 5B). Herein, genetically engineered Synechococcus elongatus PCC7942 underwent expression with several key enzymes; i.e., Rubisco, acetolactate synthase (AlsS), aceto-hydroxy acid isomeroreductase (IlvC), dihydroxy-acid dehydratase (IlvD) and 2-ketoacid decarboxylase (Kdc). Since Rubisco was the bottleneck in carbon fixation in Cyanobacteria, overexpression of this enzyme could increase the reaction rate of CO2 and, thus, the productivity rate of isobutyraldehyde.
Figure 4. (A) Chemical route for the attachment of enzymes and cofactor onto polystyrene particles. Reproduced with permission from El-Zahab et al. [74] Copyright 2008 Wiley. (B) GelCSi microcapsules with enzymes. Reproduced with permission from Wang et al. [57] Copyright 2014 American Chemical Society. (C) Co-immobilization and sequential immobilization of enzymes in membranes for methanol production from CO2. Reproduced with permission from Luo et al. [17] Copyright 2015 Elsevier. (D) Electrochemical CO2 reduction using enzymes. Reproduced with permission from Schlager et al. [75] Copyright 2016 Wiley. (E) CO2-methanol cycle using immobilized FDH, FaDH, and ADH. Originally published in Marques Netto et al. [77] under Creative Commons Attribution 4.0 International License.
Figure 4. (A) Chemical route for the attachment of enzymes and cofactor onto polystyrene particles. Reproduced with permission from El-Zahab et al. [74] Copyright 2008 Wiley. (B) GelCSi microcapsules with enzymes. Reproduced with permission from Wang et al. [57] Copyright 2014 American Chemical Society. (C) Co-immobilization and sequential immobilization of enzymes in membranes for methanol production from CO2. Reproduced with permission from Luo et al. [17] Copyright 2015 Elsevier. (D) Electrochemical CO2 reduction using enzymes. Reproduced with permission from Schlager et al. [75] Copyright 2016 Wiley. (E) CO2-methanol cycle using immobilized FDH, FaDH, and ADH. Originally published in Marques Netto et al. [77] under Creative Commons Attribution 4.0 International License.
Catalysts 13 00611 g004

4. Applications of Chemo-Enzymatic Approaches for CO2 Upcycling

Elevated average worldwide temperatures, also known as “global warming” and “climate change,” have sparked a global movement to cut the quantity of carbon dioxide (CO2) in the atmosphere. Popular strategies that have been considered and used to reduce CO2 emissions include CO2 capture and storage (CCS) and turning CO2 into fuels and chemicals. The CO2 Utilization Program of the US Department of Energy (DOE) principally emphasizes the mineralization and polymerization processes, CO2 improved retrieval of hydrocarbons, and production of chemicals [80]. These industries have a significant chance of converting high CO2 capabilities into cost-effective products. Despite the many products that can be made from CO2, there is a significant difference between the amount of CO2 produced and the amount of CO2 transformed. Using CO2 to generate fuels and chemicals is a win-win method for reducing atmospheric CO2 and effectively utilizing carbon resources [10], as illustrated in Figure 6. Many scientists are eager to investigate and create new catalytic methods that will allow CO2 to be converted into platform chemicals that may be used to create bio-based products. This section will explore various applications of the upcycling of CO2 using a chemo-enzymatic approach.
The technique, known as chemo-enzymatic, integrates biocatalytic reaction with the production of bulk chemicals. The enzymatic approach offers various benefits, including greater selectivity, yield, and the lack of changeable reaction conditions. The production of such chemicals with excellent stereoselectivity is crucial in the current environment [81]. Diverse methods or strategies have been documented for the systematic synthesis of enzyme-based catalysts, for optimizing reaction processes, and for the clarification of reaction mechanisms in order to improve the effectiveness of the enzyme catalytic process. Additionally, carbon capture and storage (CCS) technology has gained widespread adoption to remove and recycle CO2. One notable advancement in CCS technology is carbon capture, utilization, and storage (CCUS), which has the potential to reduce CO2 emissions substantially. The captured CO2 can be directed towards a novel output process, where it undergoes transformation and is utilized in a variety of ways [82]. This procedure efficiently accomplishes the upcycling and reusing of CO2 and goes beyond simple storage. Table 2 summarizes the products formed via the chemo-enzymatic approach of utilizing CO2. The advantages and disadvantages of CO2 conversion via chemical (electrochemical and catalytic thermal) and chemo-enzymatic approaches are summarized in Table 3.
Figure 6. Sustainable CO2 conversion and utilization.
Figure 6. Sustainable CO2 conversion and utilization.
Catalysts 13 00611 g006

4.1. Conversion into Minerals

In order to speed up the hydration and dehydration of carbon dioxide, carbonic anhydrase (CA) is a crucial metalloenzyme in biological systems. CA is mainly responsible for the (inter)conversion between CO2 and bicarbonate. It is known that the enzyme carbonic anhydrase increases CO2 absorption rates. A divalent zinc ion in the enzyme’s active site catalyzes the process. With reaction speeds of up to 106 s−1, CA is one of the fastest recorded enzymatic reactions [96]. CA-catalyzed processes have been shown effective in capturing gaseous carbon dioxide from industrial emissions in aqueous solutions. The three most widely used techniques for CA-catalytic conversion of CO2, including capture, sequestration, and utilization, are absorption, membrane separation, and precipitation/mineralization. These methods offer promising solutions for reducing carbon emissions in industrial settings [64]. The chemical absorption method, which uses CA-catalytic CO2 conversion with regenerable alkaline aqueous solvents, is one of the most prominent applications of the chemo-enzymatic process of CO2 conversion.
The biogenic precipitation of calcium carbonate, an important natural process, is facilitated by CA. As a result, biomimetic techniques have great potential for CO2 capture, storage, and utilization. The first step in CO2 sequestration is separating CO2 from other gaseous components in flue gas produced during fuel combustion. A common chemical absorption method is utilized, whereby CO2 is removed from flue gas in an absorber column and desorbed in a heated stripper column to yield relatively pure CO2 (Figure 7A). By utilizing immobilized CA as a packed column membrane contractor, these scrubbers can emit CO2 gas and precipitate CaCO3 (Figure 7B). Furthermore, coupling a CA-catalyzed conversion process with a chemical absorption process may increase CO2 absorption rates in alkaline aqueous solvents while also enabling the use of aqueous solvents with reduced heat/energy input for desorption.
CO2 Solution Inc. has developed a commercialized technology for biomimetic CO2 capture and storage. It employs a patented biotechnology infrastructure that can be used in the coal-fired power generating industry, the oil sands industry, as well as other CO2-intensive industries, using CA as a biocatalyst. Cement plants release CO2, which CO2 Solution Inc. collects and transforms into bicarbonate ions, which are then transformed into limestone and utilized as raw goods for cement production [97], as shown in Figure 7B. Calcite is the main byproduct of CaCO3 catalyzed by CA. The calcite CaCO3 minerals are currently employed in numerous industrial processes. These consist of a medicinal source in antacids and calcium supplements, cement constituent, paper filler and coating, white coloring in paints, and granulating excipient. Calcite is also a component of limestone, marble, and chalk [98].
Despite the numerous advantages of the chemical absorption process, it has been observed that harsh conditions, including high temperatures, high levels of organic amines, and trace pollutants, can result in enzyme denaturation. This is a major limitation in the industrial process. Various methods have been employed to overcome this limitation in investigating enzyme-enhanced CO2 sequestration. These methods include the purification and analysis of CA from different sources [99,100], as well as the immobilization of CA to enhance its resistance to abrasive environments [101]. Many studies have been done exploring the immobilization technique to improve the CA for high CO2 sequestration, which includes using materials such as mesoporous silica [102], nanofibers [103], nanoparticles [104], and membranes [105]. By co-precipitating enzymes and inorganic components, several advanced methods for enzyme immobilization have recently been devised. Recently, novel CA from alkaliphilic cyanobacterium showed improved thermal stability and outstanding CO2 collection capabilities [106]. Another recent discovery in improving CO2 sequestration technology and linking it to additional reactions for upcycling is the acceleration of CO2 sequestration via pressure. This is done by introducing carbonic anhydrase into a high-pressure reactor to catalyze the hydration of CO2 and produces precipitated CaCO3 [107]. All these efforts are made to enhance the sequestration, upcycling and recycling of CO2.
Recent advancements in metabolic and protein engineering have paved the way for using synthetic biology as a powerful tool for enhancing biological function and commercial viability. An example of this can be seen in the successful expression of Methanobacterium thermoautotrophicum in the periplasm of E. coli, creating a whole-cell biocatalyst for CO2 hydration. These processes have been shown to effectively hydrate CO2 with catalytic yield and efficacy comparable to free enzymes. However, a lower yield disparity may result from mass transportation restrictions. In comparing the operation with and without calcium components (carbonate and chloride), it has been observed that the presence of CaCl2 and CaCO3 results in a 50–70% increase in carbon precipitation [108].
The latest development in 3D printing has enabled the creation of a novel apparatus consisting of buoyant and biocatalytic constituents, which has enhanced interfacial CA catalysis to utilize and convert CO2 into CaCO3 effectively. This innovation has simplified the alignment and positioning of immobilized CA at the interfacial region. The resulting bioreactor has demonstrated the ability to expedite CO2 hydration at gas-liquid interfaces, thereby expediting the conversion of CO2 [109].

4.2. Conversion into Chemicals

The economic worth of CO2 can be increased by using it to synthesize bulk compounds. The most prominent byproducts for CO2 utilization in conventional chemical industry operations are urea and salicylic acid, and urea generation is the most significant chemical route for CO2 utilization [110]. The interaction of ammonia with CO2 is the primary industrial technique for producing urea. The production of methylammonium using supercritical NH3 and CO2 is the first reaction in the urea manufacturing process. The second reaction is the dehydration of methylammonium to produce urea. An efficient catalyst for the concurrent electrocatalytic coupling of N2 and CO2 to accomplish sustainable urea production was discovered to be PdCu alloy nanoparticles on TiO2 nanosheets [91]. In order to accelerate the synthesis of urea, *N—N* spontaneously established a C-N bond with the CO created by reducing CO2. Due to its superior stereo-specificity and region/chemo-selectivity, the enzymatic method—which was inspired by the CO2 metabolic process in cells—offers a green and effective substitute for effective CO2 conversion. The decarboxylase enzyme is one of the popular enzymes used to convert CO2 into biodegradable chemicals.
An example of this is using pyruvate decarboxylase in biodegradable chemical conversion, which offers the advantage of an environmentally safe technique [48]. These decarboxylases have drawn a lot of interest recently because they can also chemically catalyze the carboxylation of the substrate with CO2. Four distinct enzymatic carboxylation reactions obtained from catabolic pathways have been created in vitro, as stated by Faber and co-workers: (1) the carboxylation of epoxides, (2) the carboxylation of aromatics, (3) the carboxylation of hetero-aromatics, and (4) the carboxylation of aliphatics [64].
CO2 with phenol can produce salicylic acid under high pressure and temperature. A study by Ijima discovered that supercritical CO2 and phenol would successfully synthesize salicylic acid at room temperature in the presence of a Lewis acid catalyst. Under specific circumstances, the phenol was transformed into salicylic acid by adding 10 mmol of aluminium bromide [111]. In an alternative investigation, the catalytic activities of various alkali metals were evaluated. The findings indicate that potassium carbonate positively affected the production of salicylic acid from supercritical CO2 and phenol. The results demonstrated that potassium carbonate was the most efficient catalyst among all the base catalysts, with a 99% selectivity for salicylic acid carboxylation. The key impact was achieved by optimizing the reaction parameters to maintain reaction efficiency. Under specific conditions, the conversion rate of phenol was 69.3%, while salicylic acid’s production and selectivity reached 69.3% and 99%, respectively. These results suggest that potassium carbonate has great potential as a catalyst for the carboxylation of phenol to produce salicylic acid [112].

4.3. Conversion into Organic Fuels

Catalytic hydrogenation can be utilized to convert CO2 into organic compounds such as methane, methanol, formic acid, and dimethyl ether. This process produces “green fuel” that, unlike traditional fossil fuels, does not generate additional CO2 emissions upon consumption.

4.3.1. Conversion into Methane and Methanol

Methane, the principal constituent of natural gas, is a critical energy source for residential and industrial use. The methanation process, which involves converting a mixture of CO2 and H2 gases with an H2 to CO2 ratio of 4, is responsible for the production of CH4. This exothermic reaction requires a catalyst and heating for efficient conversion. Extensive research has identified Co, Ni, Ru, and other active elements as effective catalysts for this reaction [113]. Martins et al. developed a novel adsorption-conversion cycle technology focused on CO2 hydrogenation that allowed for the simultaneous isolation and use of CO2 [114]. In this method, a concurrent adsorption reactor isolates the biogas well (a CO2/CH4 combination), and it also produces methane by reacting the separated CO2 with the entering H2. This is a very promising approach for capturing and using CO2, but the advancement of this study field depends on improving process operating parameters.
The conversion of carbon dioxide (CO2) to methanol through a multienzyme process has gained widespread attention as a promising pathway due to recycling the “greenhouse” gas and creating sustainable biofuels. Although methanol can be synthesized via the catalytic hydrogenation of CO2, the occurrence of the reverse water gas shift reaction (RWGS) can limit its production. This reaction consumes a portion of the hydrogen (H2), decreasing methanol yield. Simultaneously water production can also negatively impact the catalyst’s catalytic activity, compounding the reduction in methanol production [115]. Therefore, the choice of catalyst is crucial to the hydrogenation of CO2 to methanol. Li et al. investigated the Cu/ZrO2 catalyst’s catalytic mechanism for CO2 hydrogenation to methanol [116]. Because of the distinctive crystalline structure of ZrO2, it was discovered that it could react with Cu. This enhanced Cu’s dispersion stabilized Cu nanoparticles during catalytic CO2 hydrogenation, successfully prevented the RWGS reaction, and encouraged methanol production.
The multienzyme reaction creates liquid methanol, which is simpler to ship and has much greater energy capabilities. For the first time, Yoneyama et al. reported the successful electrochemical reduction of CO2 to methanol using pyrroloquinolinequinone (PQQ) as a coenzyme and FateDH and methanol dehydrogenase (MDH) as catalysts. This method provided a simple means of producing methanol straight from CO2 under benign circumstances [64]. Dave and co-workers reported a strategy involving the sequential reduction of CO2 to methanol, with multiple dehydrogenases as the catalysts and NADH as a cofactor, to benefit the essence that dehydrogenases can efficiently catalyze the reduction of CO2 in the existence of an appropriate cofactor [13].

4.3.2. Conversion into Ethanol

The potential of microbe-based ethanol generation has garnered significant attention in recent years. Conventionally, biofuels are produced through the solventogenic conversion of biomass obtained from photosynthetic microbes. However, this technique has been demonstrated to be ineffective owing to its lower energy conversion efficiency [117]. As a result, current research is focused on the direct conversion of CO2 into biofuels by photosynthetic bacteria. While only a limited number of cyanobacterial species are capable of producing ethanol as a byproduct during biological fermentation at lower concentrations, efforts are underway to increase the generation levels to make the process more cost-effective. By inserting the pyruvate decarboxylase (pdc) and alcohol dehydrogenase II (adh) genes from Z. mobilis into the chromosomes of Synechocuystis sp. PCC6803, Dexter and Fu (2009) attempted to produce ethanol and successfully produced 550 mg/L [118].

4.3.3. Conversion into Formic Acid

Formic acid is a crucial industrial compound with various applications in animal feed preservation, textile tanning, and chemical synthesis. The current global annual production of formic acid stands at 800,000 T, primarily generated by reacting methanol and carbon monoxide with a robust base [119]. Numerous studies on CO2 hydrogenation have explored various reaction media, including organic solvents, supercritical CO2, aqueous solutions, and the addition of organic or inorganic bases or task-specific ionic liquids [120]. The most extensively studied catalysts for CO2 hydrogenation to formic acid conversion are homogeneous catalysts, which offer milder reaction conditions and improved catalytic efficiency compared to heterogeneous catalysts. Catalysts that dissolve in water are often employed in CO2-to-formic acid conversion systems. According to Zhao et al., RhCl(mtppms)3 was utilized as a catalyst to hydrogenate CO2 into formic acid in a sodium formate solution [120]. In a recent study, researchers have conducted the biocatalytic conversion of CO2 into formic acid using a biocatalytic micromixer consisting of two enzymes, carbonic anhydrase (CA) and formate dehydrogenase (FDH). These two enzymes were biomineralized in a zeolitic imidazolate framework-8 composite thin film and modified with polydopamine/polyethyleneimine [121].

4.3.4. Conversion into Dimethyl Ether (DME)

The potential utilization of dimethyl ether (DME) as a substitute fuel or precursor for creating various value-added products, such as gasoline, aromatics, and olefins, has led to significant interest in converting CO2 to DME. While DME has primarily been used as a replacement for ozone-depleting chlorofluorocarbon (CFC) chemicals, such as aerosol propellant in spray bottles, there has been a growing focus on its potential as an environmentally friendly alternative fuel [122]. DME can be generated directly by CO2 hydrogenation or by dehydrating methanol after CO2 hydrogenation. Since methanol dehydration can produce DME, the catalyst that converts CO2 into dimethyl ether is typically a mixed catalyst made of catalysts for methanol synthesis and dehydration, such as a molecular sieve [123]. Many studies have recently been done to create a CO2 hydrogenation catalyst that is extremely effective [124,125]. Several approaches have been put forth to develop a catalyst with good CO2 conversion, DME selectivity, and stability properties that can make DME through one-pot hydrogenation of CO2 [126]. A more appealing substitute for the two-step process has recently been suggested: direct synthesis in a single step. The one-step procedure combines the two reactions of methanol synthesis (via CO hydrogenation) and methanol dehydration to DME in the same reactor while utilizing hybrid bifunctional catalysts in environments that encourage methanol synthesis. The hybrid catalysts required for direct DME production need to combine acid sites for subsequent methanol dehydration and metal sites for selective hydrogenation of CO to methanol [127]. Some elements of the bifunctional catalyst, the Cu-based catalyst, are anticipated to continue being the most effective catalyst for the CO2-to-methanol reaction step.

4.3.5. Conversion into Olefins

Light olefins (C2=C4=), which are required to manufacture high value-added substances and plastics, are in great supply on a global scale. Processing hydrocarbons, the direct/indirect conversion of synthesis gas (CO + H2), and the hydrogenation of CO2 can all be used to create light olefins. The most recent technique to be researched among these is the catalytic hydrogenation of CO2, which may help reduce atmospheric CO2 emissions.
The production of olefins from CO2 can take place in many distinct ways (Figure 8), based on the catalyst being utilized: (1) the CO-mediated altered Fischer-Tropsch to olefins (FTO) route; (2) the methanol to olefins (MTO) route [27]; and (3) the direct CO2 hydrogenation to olefins utilizing promoted and multifunctional catalysts (Equation (6)).
nCO2 + 3nH2 → CnH2n + 2nH2O
It was reported that the FTO route (1) is much more effective than the MTO technique in the context of hydrocarbon generation [128]. Therefore, to optimize the production of olefins through the methanol-to-olefin (MTO) route, minimizing the formation of byproducts such as CO, CH4, C2-C4 alkanes and C5+ hydrocarbons is crucial. Significant CO2 conversions can encourage the production of such byproducts, underscoring the need to mitigate their generation. Conventionally, the MTO process involves two separate reactors: the first for methanol production and the second for olefin production over zeolite catalysts. However, a more efficient approach involves employing a bifunctional catalyst with metallic and acidic properties within a single vessel for the catalyzed reaction [129].
A recent study conducted by Guo et al. developed a fully integrated catalyst with an enzyme-like structure using calcined CC ash, iron carbides, and alkali boosters to convert CO2 hydrogenation directly into linear α-olefins (LAO), with selectivity in hydrocarbons of at least 40% and a total olefin selectivity in hydrocarbons of 72%. The output specificity in this “CO2-LAO” process came close to Fischer–Tropsch synthesis (FTS). It is shown that the bio-promoter reconfigured iron catalysts have a synergistic impact on enhancing CO2 hydrogenation efficiency in comparison to unpromoted and chemically supported iron-based catalysts [130].

4.4. CO2 Utilization in Polymer Synthesis

In the area of CO2 chemical upcycling, creating CO2 polymers with a commercial benefit from CO2 is a major field of study. The two most viable ways of CO2 upcycling for the creation of synthetic polymers are the generation of polycarbonate (PC) and polyurethane (PU) from CO2 [131].

4.4.1. Chemo-Enzymatic Synthesis of Polycarbonates

The thermoplastic polymer containing carbonate units, commonly known as polycarbonate (PC), has attracted significant attention owing to its potential for recyclability. Notably, PC is biodegradable and can be synthesized using CO2 as a comonomer. The current state-of-the-art approach for producing aliphatic polycarbonate (PPC) involves copolymerizing CO2 with epoxide. Extensive research has focused on developing new catalysts and catalytic mechanisms to enhance PPC production. This innovation is crucial for meeting the increasing demand for sustainable and environmentally friendly plastics. Liu has recently reported the discovery of a novel bimetallic nickel compound, which exhibited exceptional catalytic activity in accelerating the copolymerization of CO2 and epoxides. This finding has led to the successful synthesis of CO2-based poly(cyclohexene carbonate)s (PCHCs) with carbonate linkages exceeding 99%, via the copolymerization of CO2 and cyclohexene oxide, and under the catalytic influence of the bimetallic nickel compound. This breakthrough finding has significant implications for developing sustainable and environmentally-friendly polymer materials [132].

4.4.2. Chemo-Enzymatic Synthesis of Polyurethanes

Polyurethane (PU) is a commonly used polymer in manufacturing various polymeric materials and plastic sponges. Due to its remarkable properties, it has found widespread use across numerous industries following years of scientific research and development. Nonetheless, the conventional synthesis methods of PU involve the utilization of isocyanates, which pose significant safety risks [133]. A new development on the path to the production of PU is the substitution of isocyanate with CO2. This will not only reduce the issue of CO2 emissions but also realize the resource exploitation of CO2. Several recent advancements in the production of PU include employing CO2 as a carbonyl source to swap out isocyanate and polysiloxane with amino modifications [134], synthesizing polyurethane thermoplastics via polycondensation of CO2 with mixed binary diamines [135], and the recent disclosure of a procedure for the indirect production of polyurethane from CO2 that is solvent- and catalyst-free [136].

5. Significance of Chemo-Enzymatic Methods in Comparison

Chemo-enzymatic describes a synthetic strategy that integrates biocatalytic reactions for large-scale chemical synthesis. The enzymatic approach offers multiple benefits, including enhanced selectivity, yield, and tolerance towards diverse reaction conditions [27]. In the current situation, producing such chemicals with high stereoselectivity is important [81]. A schematic representation of the chemo-enzymatic approach to CO2 conversion is presented in Figure 9.
The chemo-enzymatic route has demonstrated the ability to convert CO2 into bulk chemicals. The process involves the conversion of CO2 into CO by [NiFe] carbon monoxide dehydrogenase, which can then be further transformed into various value-added products, such as hydrocarbons, methanol, and acetic acid [24]. Previous studies have extensively investigated the production of formate, formic acid, and methanol via the enzymatic reduction of CO2 [137]. Additionally, combining electrochemical and chemo-enzymatic approaches has resulted in a Faradic efficiency of 10% [75]. CO2 can also be converted into methane via the enzyme nitrogenase. Yang et al. [84] studied nitrogenase for methane production, where 1 nmol of nitrogenase produced 21 nmol of methane at optimized conditions. Carbonic anhydrase (CA) is another enzyme with a higher reaction rate for CO2 conversion when compared with other enzymes [27]. CA is also involved in the process of CO2 capture and CO2 sequestration. Although this reaction requires high temperature, which might affect the enzyme activity, this setback can be overcome by enzyme immobilization onto beads/matrix materials. In this regard, this process further assists CO2 capture and conversion in producing valuable products [138]. One notable chemo-enzymatic mechanism is the Michael addition, where the enzyme lyase serves as the catalyst. The carboxylation of molecules, involving the cleavage of chemical bonds and leading to the formation of valuable chemical compounds, can be accomplished by utilizing this enzyme [64].
The chemo-enzymatic approach towards CO2 conversion is a promising avenue for achieving high selectivity and product yield. However, there are some potential drawbacks associated with this approach, including high cost; and efforts have been made to increase its activity and stability. It is crucial to conduct scale-up pilot research studies in industrial settings to assess the viability of this approach for large-scale applications. Moreover, computational methods can be employed to design novel enzymes that can enhance the efficiency of CO2 conversion. These advances in enzyme design hold significant potential for achieving higher yields and reducing the overall cost of the process. The development of new inventions and research in the field of CO2 capture, utilization, and sequestration is gaining momentum worldwide. The goal is to move towards sustainable solutions that can mitigate the impact of CO2 emissions on the environment. Researchers, governments, and people are working together to promote this vital research, which is expected to play a crucial role in achieving global sustainability targets.
While turning waste CO2 into high-value products is a promising complementary strategy for improving the carbon balance and financial prospects. Bioenergy transformation to carbon capture and storage (BECCS) is also gaining considerable attention as a potential carbon-negative technology for the future. Combining chemo- and bio-catalysis can produce beneficial synergies in the chemical value chain when one of the two catalytic fields is insufficient. The long-standing research goal of combining these two catalytic fields represents an economically and environmentally appealing concept, as it can potentially avoid time- and capacity-consuming work-up stages of intermediates.

6. Conclusions

The increasing atmospheric CO2 concentration is a significant environmental issue, leading to global warming. In response, efforts are being made to collect and transform CO2 into value-added goods. CO2 utilization can be directly used or transformed into compounds with added value. Enzymatic technology is a promising approach, utilizing CO2 metabolic processes in cells and providing an environmentally friendly way to upcycle CO2. Many methods for rational enzyme-based catalyst synthesis, reaction process refinement, and reaction mechanism clarification have been documented, contributing to improved chemo-enzyme catalytic activity. Despite recent advancements, more research is necessary for large-scale utilization, cheaper and environmentally stable catalysts, and the industrialization of developed technologies. Finding low-energy pathways for CO2 conversion through efficient enzyme-based technology requires further exploration. The significant scientific and technical advances in CO2 conversion and upcycling presented in this review provide a basis for future research to address the urgent need for effective CO2 utilization and the reduction of carbon emissions.

Author Contributions

Conceptualization, M.K. and K.M.; methodology, M.K. and K.M.; formal analysis, K.M., R.S. and Y.W.T.; investigation, K.M., R.S. and Y.W.T.; resources, H.S., S.R. and S.M.; data curation, K.M., R.S. and Y.W.T.; writing—original draft preparation, K.M., R.S. and Y.W.T.; writing—review and editing, K.M., S.M., H.S. and M.K.; visualization, R.S. and Y.W.T.; supervision, M.K. and K.M.; project administration, M.K. and K.M.; funding acquisition, M.K. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thank Sunway University for the International Research Networks Grant Scheme 2.0 (STR-IRNGS-SET-GAMRG-01-2022) for this work.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Najafabadi, A.T. CO2 chemical conversion to useful products: An engineering insight to the latest advances toward sustainability. Int. J. Energy Res. 2013, 37, 485–499. [Google Scholar] [CrossRef]
  2. Van Calster, G.; Reins, L. The Paris Agreement on Climate Change. In Paris Agreement on Climate Change: A Commentary; Edward Elgar Publishing: Cheltenham, UK, 2021; pp. 11–22. [Google Scholar] [CrossRef]
  3. Krzywanski, J.; Ashraf, W.M.; Czakiert, T.; Sosnowski, M.; Grabowska, K.; Zylka, A.; Kulakowska, A.; Skrobek, D.; Mistal, S.; Gao, Y. CO2 Capture by Virgin Ivy Plants Growing up on the External Covers of Houses as a Rapid Complementary Route to Achieve Global GHG Reduction Targets. Energies 2022, 15, 1683. [Google Scholar] [CrossRef]
  4. International Energy Agency. An Energy Sector Roadmap to Carbon Neutrality in China, An Energy Sect. Roadmap to Carbon Neutrality China; IEA: Paris, France, 2021. [Google Scholar] [CrossRef]
  5. Markandan, K.; Chai, W.S. Perspectives on Nanomaterials and Nanotechnology for Sustainable Bioenergy Generation. Materials 2022, 15, 7769. [Google Scholar] [CrossRef]
  6. Alshalif, A.F.; Irwan, J.M.; Othman, N.; Al-Gheethi, A.A.; Shamsudin, S. A systematic review on bio-sequestration of carbon dioxide in bio-concrete systems: A future direction. Eur. J. Environ. Civ. Eng. 2022, 26, 1209–1228. [Google Scholar] [CrossRef]
  7. Sharma, T.; Sharma, S.; Kamyab, H.; Kumar, A. Energizing the CO2 utilization by chemo-enzymatic approaches and potentiality of carbonic anhydrases: A review. J. Clean. Prod. 2020, 247, 119138. [Google Scholar] [CrossRef]
  8. Kumaravel, V.; Bartlett, J.; Pillai, S.C. Photoelectrochemical Conversion of Carbon Dioxide (CO2) into Fuels and Value-Added Products. ACS Energy Lett. 2020, 5, 486–519. [Google Scholar] [CrossRef] [Green Version]
  9. Mets, L. System for the Production of Methane from CO2. European Patent EP2032709A1, 3 November 2009. [Google Scholar]
  10. Van Duc Long, N.; Lee, J.; Koo, K.K.; Luis, P.; Lee, M. Recent progress and novel applications in enzymatic conversion of carbon dioxide. Energies 2017, 10, 473. [Google Scholar] [CrossRef] [Green Version]
  11. Alpdağtaş, S.; Turunen, O.; Valjakka, J.; Binay, B. The challenges of using NAD+-dependent formate dehydrogenases for CO2 conversion. Crit. Rev. Biotechnol. 2022, 42, 953–972. [Google Scholar] [CrossRef]
  12. Han, G.H.; Bang, J.; Park, G.; Choe, S.; Jang, Y.J.; Jang, H.W.; Kim, S.Y.; Ahn, S.H. Recent Advances in Electrochemical, Photochemical, and Photoelectrochemical Reduction of CO2 to C2+ Products. Small 2023, 2205765. [Google Scholar] [CrossRef]
  13. Obert, R.; Dave, B.C. Enzymatic conversion of carbon dioxide to methanol: Enhanced methanol production in silica sol-gel matrices. J. Am. Chem. Soc. 1999, 121, 12192–12193. [Google Scholar] [CrossRef]
  14. Baskaya, F.S.; Zhao, X.; Flickinger, M.C.; Wang, P. Thermodynamic feasibility of enzymatic reduction of carbon dioxide to methanol. Appl. Biochem. Biotechnol. 2010, 162, 391–398. [Google Scholar] [CrossRef]
  15. Cazelles, R.; Drone, J.; Fajula, F.; Ersen, O.; Moldovan, S.; Galarneau, A. Reduction of CO2 to methanol by a polyenzymatic system encapsulated in phospholipids-silica nanocapsules. New J. Chem. 2013, 37, 3721–3730. [Google Scholar] [CrossRef]
  16. Marpani, F.; Pinelo, M.; Meyer, A.S. Enzymatic conversion of CO2 to CH3OH via reverse dehydrogenase cascade biocatalysis: Quantitative comparison of efficiencies of immobilized enzyme systems. Biochem. Eng. J. 2017, 127, 217–228. [Google Scholar] [CrossRef] [Green Version]
  17. Luo, J.; Meyer, A.S.; Mateiu, R.V.; Pinelo, M. Cascade catalysis in membranes with enzyme immobilization for multi-enzymatic conversion of CO2 to methanol. New Biotechnol. 2015, 32, 319–327. [Google Scholar] [CrossRef] [PubMed]
  18. Yang, H.; Kaczur, J.J.; Sajjad, S.D.; Masel, R.I. Electrochemical conversion of CO2 to formic acid utilizing SustainionTM membranes. J. CO2 Util. 2017, 20, 208–217. [Google Scholar] [CrossRef]
  19. Hori, Y.; Wakebe, H.; Tsukamoto, T.; Koga, O. Electrocatalytic process of CO selectivity in electrochemical reduction of CO2 at metal electrodes in aqueous media. Electrochim. Acta 1994, 39, 1833–1839. [Google Scholar] [CrossRef]
  20. Malkhandi, S.; Yeo, B.S. Electrochemical conversion of carbon dioxide to high value chemicals using gas-diffusion electrodes. Curr. Opin. Chem. Eng. 2019, 26, 112–121. [Google Scholar] [CrossRef]
  21. Yuan, J.; Hao, C. Solar-driven photoelectrochemical reduction of carbon dioxide to methanol at CuInS2 thin film photocathode. Sol. Energy Mater. Sol. Cells 2013, 108, 170–174. [Google Scholar] [CrossRef]
  22. Yeo, J.S.; Lee, J.H.; Yoo, E.J. Electrochemical properties of environment-friendly lithium-tin liquid metal battery. Electrochim. Acta 2018, 290, 228–235. [Google Scholar] [CrossRef]
  23. Rosen, B.A.; Salehi-Khojin, A.; Thorson, M.R.; Zhu, W.; Whipple, D.T.; Kenis, P.J.A.; Masel, R.I. Ionic Liquid—Mediated Selective Conversion of CO2 to CO at Low Overpotentials. Science 2011, 334, 643–644. [Google Scholar] [CrossRef]
  24. Parkin, A.; Seravalli, J.; Vincent, K.A.; Ragsdale, S.W.; Armstrong, F.A. Rapid and efficient electrocatalytic CO2/CO interconversions by Carboxydothermus hydrogenoformans CO dehydrogenase I on an electrode. J. Am. Chem. Soc. 2007, 129, 10328–10329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Shin, W.; Lee, S.H.; Shin, J.W.; Lee, S.P.; Kim, Y. Highly Selective Electrocatalytic Conversion of CO2 to CO at −0.57 V (NHE) by Carbon Monoxide Dehydrogenase from Moorella thermoacetica. J. Am. Chem. Soc. 2003, 125, 14688–14689. [Google Scholar] [CrossRef] [PubMed]
  26. Zarandi, R.F.; Rezaei, B.; Ghaziaskar, H.S.; Ensafi, A.A. Electrochemical conversion of CO2 to methanol using a glassy carbon electrode, modified by Pt@histamine-reduced graphene oxide. Int. J. Hydrogen Energy 2019, 44, 30820–30831. [Google Scholar] [CrossRef]
  27. Saravanan, A.; Kumar, P.S.; Vo, D.V.N.; Jeevanantham, S.; Bhuvaneswari, V.; Narayanan, V.A.; Yaashikaa, P.R.; Swetha, S.; Reshma, B. A comprehensive review on different approaches for CO2 utilization and conversion pathways. Chem. Eng. Sci. 2021, 236, 116515. [Google Scholar] [CrossRef]
  28. Huang, C.; Mao, D.; Guo, X.; Yu, J. Microwave-Assisted Hydrothermal Synthesis of CuO–ZnO–ZrO2 as Catalyst for Direct Synthesis of Methanol by Carbon Dioxide Hydrogenation. Energy Technol. 2017, 5, 2100–2107. [Google Scholar] [CrossRef]
  29. Ud, I.; Shaharun, M.S.; Naeem, A.; Tasleem, S.; Ra, M. Carbon nano fiber-based copper/zirconia catalyst for hydrogenation of CO2 to methanol. J. CO2 Util. 2017, 21, 145–155. [Google Scholar]
  30. Liang, Z.; Gao, P.; Tang, Z.; Lv, M.; Sun, Y. Three dimensional porous Cu-Zn/Al foam monolithic catalyst for CO2 hydrogenation to methanol in microreactor. J. CO2 Util. 2017, 21, 191–199. [Google Scholar] [CrossRef]
  31. Collins, S.E.; Baltanás, M.A.; Bonivardi, A.L. An infrared study of the intermediates of methanol synthesis from carbon dioxide over Pd/β-Ga2O3. J. Catal. 2004, 226, 410–421. [Google Scholar] [CrossRef]
  32. Bonivardi, A.L.; Chiavassa, D.L.; Querini, C.A.; Baltanás, M.A. Enhancement of the catalytic performance to methanol synthesis from CO2/H2 by gallium addition to palladium/silica catalysts. Stud. Surf. Sci. Catal. 2000, 130, 3747–3752. [Google Scholar] [CrossRef]
  33. Dorner, R.W.; Hardy, D.R.; Williams, F.W.; Willauer, H.D. Effects of ceria-doping on a CO2 hydrogenation iron-manganese catalyst. Catal. Commun. 2010, 11, 816–819. [Google Scholar] [CrossRef]
  34. Galadima, A.; Muraza, O. Catalytic thermal conversion of CO2 into fuels: Perspective and challenges. Renew. Sustain. Energy Rev. 2019, 115, 109333. [Google Scholar] [CrossRef]
  35. Ning, W.; Koizumi, N.; Yamada, M. Researching Fe catalyst suitable for CO2-containing syngas for Fischer-Tropsch synthesis. Energy Fuels 2009, 23, 4696–4700. [Google Scholar] [CrossRef]
  36. Visconti, C.G.; Martinelli, M.; Falbo, L.; Fratalocchi, L.; Lietti, L. CO2 hydrogenation to hydrocarbons over Co and Fe-based Fischer-Tropsch catalysts. Catal. Today 2016, 277, 161–170. [Google Scholar] [CrossRef]
  37. Rodriguez, J.A.; Evans, J.; Feria, L.; Vidal, A.B.; Liu, P.; Nakamura, K. CO2 hydrogenation on Au/TiC, Cu/TiC, and Ni/TiC catalysts: Production of CO, methanol, and methane. J. Catal. 2013, 307, 162–169. [Google Scholar] [CrossRef]
  38. Asara, G.G.; Ricart, J.M.; Rodriguez, J.A.; Illas, F. Exploring the activity of a novel Au/TiC(001) model catalyst towards CO and CO2 hydrogenation. Surf. Sci. 2015, 640, 141–149. [Google Scholar] [CrossRef] [Green Version]
  39. Li, Y.; Liu, G.; Kong, W.; Zhang, S.; Bao, Y.; Zhao, H.; Wang, L.; Zhou, L.; Jiang, Y. Electrocatalytic NAD(P)H regeneration for biosynthesis. Green Chem. Eng. 2023. [Google Scholar] [CrossRef]
  40. Yan, L.; Liu, G.; Liu, J.; Bai, J.; Li, Y.; Chen, H.; Zhou, L.; Gao, J.; Jiang, Y. Hierarchically porous metal organic framework immobilized formate dehydrogenase for enzyme electrocatalytic CO2 reduction. Chem. Eng. J. 2022, 450, 138164. [Google Scholar] [CrossRef]
  41. Reda, T.; Plugge, C.M.; Abram, N.J.; Hirst, J. Reversible interconversion of carbon dioxide and formate by an electroactive enzyme. Proc. Natl. Acad. Sci. USA 2008, 105, 10654–10658. [Google Scholar] [CrossRef] [Green Version]
  42. Beller, M.; Bornscheuer, U.T. CO2 Fixation through Hydrogenation by Chemical or Enzymatic Methods. Angew. Chem. Int. Ed. 2014, 53, 4527–4528. [Google Scholar] [CrossRef] [PubMed]
  43. Schuchmann, K.; Müller, V. Direct and reversible hydrogenation of CO2 to formate by a bacterial carbon dioxide reductase. Science 2013, 342, 1382–1385. [Google Scholar] [CrossRef] [PubMed]
  44. Hull, J.F.; Himeda, Y.; Wang, W.H.; Hashiguchi, B.; Periana, R.; Szalda, D.J.; Muckerman, J.T.; Fujita, E. Reversible hydrogen storage using CO2 and a proton-switchable iridium catalyst in aqueous media under mild temperatures and pressures. Nat. Chem. 2012, 4, 383–388. [Google Scholar] [CrossRef] [PubMed]
  45. Bachmeier, A.; Wang, V.C.C.; Woolerton, T.W.; Bell, S.; Fontecilla-Camps, J.C.; Can, M.; Ragsdale, S.W.; Chaudhary, Y.S.; Armstrong, F.A. How Light-Harvesting Semiconductors Can Alter the Bias of Reversible Electrocatalysts in Favor of H2 Production and CO2 Reduction. J. Am. Chem. Soc. 2013, 135, 15026–15032. [Google Scholar] [CrossRef] [PubMed]
  46. Woolerton, T.W.; Sheard, S.; Reisner, E.; Pierce, E.; Ragsdale, S.W.; Armstrong, F.A. Efficient and Clean Photoreduction of CO2 to CO by Enzyme-Modified TiO2 Nanoparticles Using Visible Light. J. Am. Chem. Soc. 2010, 132, 2132–2133. [Google Scholar] [CrossRef] [Green Version]
  47. Zuo, Y.; Sheng, W.; Tao, W.; Li, Z. Direct methanol fuel cells system—A review of dual-role electrocatalysts for oxygen reduction and methanol oxidation. J. Mater. Sci. Technol. 2022, 114, 29–41. [Google Scholar] [CrossRef]
  48. Ünlü, A.; Duman-Özdamar, Z.E.; Çaloğlu, B.; Binay, B. Enzymes for Efficient CO2 Conversion. Protein J. 2021, 40, 489–503. [Google Scholar] [CrossRef]
  49. Yang, Z.; Moure, V.R.; Dean, D.R.; Seefeldt, L.C. Carbon dioxide reduction to methane and coupling with acetylene to form propylene catalyzed by remodeled nitrogenase. Proc. Natl. Acad. Sci. USA 2012, 109, 19644–19648. [Google Scholar] [CrossRef] [Green Version]
  50. Zhang, Y.T.; Zhang, L.; Chen, H.L.; Zhang, H.M. Selective separation of low concentration CO2 using hydrogel immobilized CA enzyme based hollow fiber membrane reactors. Chem. Eng. Sci. 2010, 65, 3199–3207. [Google Scholar] [CrossRef]
  51. Vinoba, M.; Bhagiyalakshmi, M.; Grace, A.N.; Chu, D.H.; Nam, S.C.; Yoon, Y.; Yoon, S.H.; Jeong, S.K. CO2 absorption and sequestration as various polymorphs of CaCO3 using sterically hindered amine. Langmuir 2013, 29, 15655–15663. [Google Scholar] [CrossRef]
  52. Annunziato, G.; Angeli, A.; D’Alba, F.; Bruno, A.; Pieroni, M.; Vullo, D.; De Luca, V.; Capasso, C.; Supuran, C.T.; Costantino, G. Discovery of New Potential Anti-Infective Compounds Based on Carbonic Anhydrase Inhibitors by Rational Target-Focused Repurposing Approaches. ChemMedChem 2016, 11, 1904–1914. [Google Scholar] [CrossRef]
  53. Thauer, R.K. A fifth pathway of carbon fixation. Science 2007, 318, 1732–1733. [Google Scholar] [CrossRef]
  54. Castillo, R.; Oliva, M.; Martí, S.; Moliner, V. A theoretical study of the catalytic mechanism of formate dehydrogenase. J. Phys. Chem. B 2008, 112, 10013–10022. [Google Scholar] [CrossRef] [PubMed]
  55. Yadav, R.K.; Baeg, J.O.; Oh, G.H.; Park, N.J.; Kong, K.J.; Kim, J.; Hwang, D.W.; Biswas, S.K. A photocatalyst-enzyme coupled artificial photosynthesis system for solar energy in production of formic acid from CO2. J. Am. Chem. Soc. 2012, 134, 11455–11461. [Google Scholar] [CrossRef]
  56. Ceccaldi, P.; Schuchmann, K.; Müller, V.; Elliott, S.J. The Hydrogen Dependent CO2 reductase: The first completely co tolerant fefe-hydrogenase. Energy Environ. Sci. 2017, 10, 503–508. [Google Scholar] [CrossRef] [Green Version]
  57. Wang, X.; Li, Z.; Shi, J.; Wu, H.; Jiang, Z.; Zhang, W.; Song, X.; Ai, Q. Bioinspired approach to multienzyme cascade system construction for efficient carbon dioxide reduction. ACS Catal. 2014, 4, 962–972. [Google Scholar] [CrossRef]
  58. Dibenedetto, A.; Stufano, P.; MacYk, W.; Baran, T.; Fragale, C.; Costa, M.; Aresta, M. Hybrid technologies for an enhanced carbon recycling based on the enzymatic reduction of CO2 to methanol in water: Chemical and photochemical NADH regeneration. ChemSusChem 2012, 5, 373–378. [Google Scholar] [CrossRef]
  59. Seefeldt, L.C.; Ensign, S.A.; Rasche, M.E. Carbonyl Sulfide and Carbon Dioxide as new Substrates, and Carbon Disulfide as a new Inhibitor, of Nitrogenase. Biochemistry 1995, 34, 5382–5389. [Google Scholar] [CrossRef]
  60. Seefeldt, L.C.; Yang, Z.Y.; Lukoyanov, D.A.; Harris, D.F.; Dean, D.R.; Raugei, S.; Hoffman, B.M. Reduction of Substrates by Nitrogenases. Chem. Rev. 2020, 120, 5082–5106. [Google Scholar] [CrossRef]
  61. Wu, Z.; Nan, Y.; Zhao, Y.; Wang, X.; Huang, S.; Shi, J. Immobilization of carbonic anhydrase for facilitated CO2 capture and separation. Chin. J. Chem. Eng. 2020, 28, 2817–2831. [Google Scholar] [CrossRef]
  62. Jensen, E.L.; Clement, R.; Kosta, A.; Maberly, S.C.; Gontero, B. A new widespread subclass of carbonic anhydrase in marine phytoplankton. ISME J. 2019, 13, 2094–2106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Cuffaro, D.; Nuti, E.; Rossello, A. An overview of carbohydrate-based carbonic anhydrase inhibitors. J. Enzym. Inhib. Med. Chem. 2020, 35, 1906–1922. [Google Scholar] [CrossRef]
  64. Shi, J.; Jiang, Y.; Jiang, Z.; Wang, X.; Wang, X.; Zhang, S.; Han, P.; Yang, C. Enzymatic conversion of carbon dioxide. Chem. Soc. Rev. 2015, 44, 5981–6000. [Google Scholar] [CrossRef]
  65. Zhang, S.; Lu, H.; Lu, Y. Enhanced stability and chemical resistance of a new nanoscale biocatalyst for accelerating CO2 absorption into a carbonate solution. Environ. Sci. Technol. 2013, 47, 13882–13888. [Google Scholar] [CrossRef] [PubMed]
  66. Hwang, E.T.; Gang, H.; Chung, J.; Gu, M.B. Carbonic anhydrase assisted calcium carbonate crystalline composites as a biocatalyst. Green Chem. 2012, 14, 2216–2220. [Google Scholar] [CrossRef]
  67. Forsyth, C.; Yip, T.W.S.; Patwardhan, S.V. CO2 sequestration by enzyme immobilized onto bioinspired silica. Chem. Commun. 2013, 49, 3191–3193. [Google Scholar] [CrossRef] [PubMed]
  68. Schoffelen, S.; Van Hest, J.C.M. Multienzyme systems: Bringing enzymes together in vitro. Soft Matter 2012, 8, 1736–1746. [Google Scholar] [CrossRef]
  69. Liu, Z.; Wang, K.; Chen, Y.; Tan, T.; Nielsen, J. Third-generation biorefineries as the means to produce fuels and chemicals from CO2. Nat. Catal. 2020, 3, 274–288. [Google Scholar] [CrossRef]
  70. Alissandratos, A.; Easton, C.J. Biocatalysis for the application of CO2 as a chemical feedstock. Beilstein J. Org. Chem. 2015, 11, 2370–2387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Hwang, E.T.; Lee, S. Multienzymatic Cascade Reactions via Enzyme Complex by Immobilization. ACS Catal. 2019, 9, 4402–4425. [Google Scholar] [CrossRef]
  72. Dunn, M.F.; Niks, D.; Ngo, H.; Barends, T.R.M.; Schlichting, I. Tryptophan synthase: The workings of a channeling nanomachine. Trends Biochem. Sci. 2008, 33, 254–264. [Google Scholar] [CrossRef]
  73. Conrado, R.J.; Varner, J.D.; DeLisa, M.P. Engineering the spatial organization of metabolic enzymes: Mimicking nature’s synergy. Curr. Opin. Biotechnol. 2008, 19, 492–499. [Google Scholar] [CrossRef]
  74. El-Zahab, B.; Donnelly, D.; Wang, P. Particle-Tethered NADH for Production of Methanol from CO2 Catalyzed by Coimmobilized Enzymes. Biotechnol. Bioeng. 2008, 99, 508–514. [Google Scholar] [CrossRef] [PubMed]
  75. Schlager, S.; Dumitru, L.M.; Haberbauer, M.; Fuchsbauer, A.; Neugebauer, H.; Hiemetsberger, D.; Wagner, A.; Portenkirchner, E.; Sariciftci, N.S. Electrochemical Reduction of Carbon Dioxide to Methanol by Direct Injection of Electrons into Immobilized Enzymes on a Modified Electrode. ChemSusChem 2016, 9, 631–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Kuk, S.K.; Singh, R.K.; Nam, D.H.; Singh, R.; Lee, J.-K.; Park, C.B. Photoelectrochemical Reduction of Carbon Dioxide to Methanol through a Highly Efficient Enzyme Cascade. Angew. Chemie Int. Ed. 2017, 56, 3827–3832. [Google Scholar] [CrossRef] [PubMed]
  77. Netto, C.G.C.M.; Andrade, L.H.; Toma, H.E. Carbon dioxide/methanol conversion cycle based on cascade enzymatic reactions supported on superparamagnetic nanoparticles. An. Acad. Bras. Cienc. 2018, 90, 593–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Tong, X.; El-Zahab, B.; Zhao, X.; Liu, Y.; Wang, P. Enzymatic synthesis of L-lactic acid from carbon dioxide and ethanol with an inherent cofactor regeneration cycle. Biotechnol. Bioeng. 2011, 108, 465–469. [Google Scholar] [CrossRef]
  79. Atsumi, S.; Higashide, W.; Liao, J.C. Direct photosynthetic recycling of carbon dioxide to isobutyraldehyde. Nat. Biotechnol. 2009, 27, 1177–1180. [Google Scholar] [CrossRef]
  80. Damiani, D.; Litynski, J.T.; McIlvried, H.G.; Vikara, D.M.; Srivastava, R.D. The US department of Energy’s R&D program to reduce greenhouse gas emissions through beneficial uses of carbon dioxide. Greenh. Gases Sci. Technol. 2012, 2, 9–16. [Google Scholar] [CrossRef]
  81. Biermann, M.; Gruß, H.; Hummel, W.; Gröger, H. Guerbet Alcohols: From Processes under Harsh Conditions to Synthesis at Room Temperature under Ambient Pressure. ChemCatChem 2016, 8, 895–899. [Google Scholar] [CrossRef]
  82. Alper, E.; Orhan, O.Y. CO2 utilization: Developments in conversion processes. Petroleum 2017, 3, 109–126. [Google Scholar] [CrossRef]
  83. Xu, S.; Lu, Y.; Li, J.; Jiang, Z.; Wu, H. Efficient Conversion of CO2 to Methanol Catalyzed by Three Dehydrogenases Co-Encapsulated in an Alginate-Silica (ALG-SiO2) Hybrid Gel. Ind. Eng. Chem. Res. 2006, 45, 4567–4573. [Google Scholar] [CrossRef]
  84. Yang, S.; Peng, S.; Xu, J.; Luo, Y.; Li, D. Methane and nitrous oxide emissions from paddy field as affected by water-saving irrigation. Phys. Chem. Earth Parts A/B/C 2012, 53–54, 30–37. [Google Scholar] [CrossRef]
  85. Yong, J.K.J.; Stevens, G.W.; Caruso, F.; Kentish, S.E. The use of carbonic anhydrase to accelerate carbon dioxide capture processes. J. Chem. Technol. Biotechnol. 2015, 90, 3–10. [Google Scholar] [CrossRef]
  86. Wei, J.; Ge, Q.; Yao, R.; Wen, Z.; Fang, C.; Guo, L.; Xu, H.; Sun, J. Directly converting CO2 into a gasoline fuel. Nat. Commun. 2017, 8, 15174. [Google Scholar] [CrossRef] [Green Version]
  87. Bahruji, H.; Armstrong, R.D.; Esquius, J.R.; Jones, W.; Bowker, M.; Hutchings, G.J. Hydrogenation of CO2 to Dimethyl Ether over Brønsted Acidic PdZn Catalysts. Ind. Eng. Chem. Res. 2018, 57, 6821–6829. [Google Scholar] [CrossRef]
  88. Cai, T.; Sun, H.; Qiao, J.; Zhu, L.; Zhang, F.; Zhang, J.; Tang, Z.; Wei, X.; Yang, J.; Yuan, Q.; et al. Cell-free chemo-enzymatic starch synthesis from carbon dioxide. Science 2021, 373, 1523–1527. [Google Scholar] [CrossRef] [PubMed]
  89. Desmons, S.; Grayson-Steel, K.; Nuñez-Dallos, N.; Vendier, L.; Hurtado, J.; Clapés, P.; Fauré, R.; Dumon, C.; Bontemps, S. Enantioselective Reductive Oligomerization of Carbon Dioxide into l-Erythrulose via a Chemo-enzymatic Catalysis. J. Am. Chem. Soc. 2021, 143, 16274–16283. [Google Scholar] [CrossRef] [PubMed]
  90. Calvo-Castañera, F.; Álvarez-Rodríguez, J.; Candela, N.; Maroto-Valiente, Á. First Phenol Carboxylation with CO2 on Carbon Nanostructured C@Fe-Al2O3 Hybrids in Aqueous Media under Mild Conditions. Nanomaterials 2021, 11, 190. [Google Scholar] [CrossRef]
  91. Chen, C.; Zhu, X.; Wen, X.; Zhou, Y.; Zhou, L.; Li, H.; Tao, L.; Li, Q.; Du, S.; Liu, T.; et al. Coupling N2 and CO2 in H2O to synthesize urea under ambient conditions. Nat. Chem. 2020, 12, 717–724. [Google Scholar] [CrossRef] [PubMed]
  92. Mazari, S.A.; Hossain, N.; Basirun, W.J.; Mubarak, N.M.; Abro, R.; Sabzoi, N.; Shah, A. An overview of catalytic conversion of CO2 into fuels and chemicals using metal organic frameworks. Process Saf. Environ. Prot. 2021, 149, 67–92. [Google Scholar] [CrossRef]
  93. He, Z.; Qian, Q.; Ma, J.; Meng, Q.; Zhou, H.; Song, J.; Liu, Z.; Han, B. Water-Enhanced Synthesis of Higher Alcohols from CO2 Hydrogenation over a Pt/Co3O4 Catalyst under Milder Conditions. Angew. Chem. Int. Ed. 2016, 55, 737–741. [Google Scholar] [CrossRef]
  94. Zhu, Q. Developments on CO2-utilization technologies. Clean Energy 2019, 3, 85–100. [Google Scholar] [CrossRef] [Green Version]
  95. Hu, B.; Guild, C.; Suib, S.L. Thermal, electrochemical, and photochemical conversion of CO2 to fuels and value-added products. J. CO2 Util. 2013, 2, 64. [Google Scholar] [CrossRef]
  96. Gundersen, M.T.; Von Solms, N.; Woodley, J.M. Enzymatically assisted CO2 removal from flue-gas. Energy Procedia 2014, 63, 624–632. [Google Scholar] [CrossRef] [Green Version]
  97. Kanth, B.K.; Lee, J.; Pack, S.P. Carbonic anhydrase: Its biocatalytic mechanisms and functional properties for efficient CO2 capture process development. Eng. Life Sci. 2013, 13, 422–431. [Google Scholar] [CrossRef]
  98. Hargis, C.W.; Chen, I.A.; Devenney, M.; Fernandez, M.J.; Gilliam, R.J.; Thatcher, R.P. Calcium Carbonate Cement: A Carbon Capture, Utilization, and Storage (CCUS) Technique. Materials 2021, 14, 2709. [Google Scholar] [CrossRef] [PubMed]
  99. Chafik, A.; El Hassani, K.; Essamadi, A.; Çelik, S.Y.; Mavi, A. Efficient sequestration of carbon dioxide into calcium carbonate using a novel carbonic anhydrase purified from liver of camel (Camelus dromedarius). J. CO2 Util. 2020, 42, 101310. [Google Scholar] [CrossRef]
  100. Müller, W.E.G.; Schlossmacher, U.; Schröder, H.C.; Lieberwirth, I.; Glasser, G.; Korzhev, M.; Neufurth, M.; Wang, X. Enzyme-accelerated and structure-guided crystallization of calcium carbonate: Role of the carbonic anhydrase in the homologous system. Acta Biomater. 2014, 10, 450–462. [Google Scholar] [CrossRef]
  101. Cowan, D.A.; Fernandez-Lafuente, R. Enhancing the functional properties of thermophilic enzymes by chemical modification and immobilization. Enzym. Microb. Technol. 2011, 49, 326–346. [Google Scholar] [CrossRef]
  102. Kim, H.S.; Hong, S.-G.; Woo, K.M.; Seijas, V.T.; Kim, S.; Lee, J.; Kim, J. Precipitation-Based Nanoscale Enzyme Reactor with Improved Loading, Stability, and Mass Transfer for Enzymatic CO2 Conversion and Utilization. ACS Catal. 2018, 8, 6526–6536. [Google Scholar] [CrossRef]
  103. Effendi, S.S.W.; Chiu, C.-Y.; Chang, Y.-K.; Ng, I.-S. Crosslinked on novel nanofibers with thermophilic carbonic anhydrase for carbon dioxide sequestration. Int. J. Biol. Macromol. 2020, 152, 930–938. [Google Scholar] [CrossRef]
  104. Lim, H.K.; Kim, D.R.; Hwang, I.T. Sequestration of CO2 into CaCO3 using Carbonic Anhydrase Immobilization on Functionalized Aluminum Oxide. Appl. Biochem. Microbiol. 2019, 55, 375–379. [Google Scholar] [CrossRef]
  105. Bednár, A.; Nemestóthy, N.; Bakonyi, P.; Fülöp, L.; Zhen, G.; Lu, X.; Kobayashi, T.; Kumar, G.; Xu, K.; Bélafi-Bakó, K. Enzymatically-boosted ionic liquid gas separation membranes using carbonic anhydrase of biomass origin. Chem. Eng. J. 2016, 303, 621–626. [Google Scholar] [CrossRef]
  106. Duan, L.; Li, H.; Zhang, Y. Synthesis of Hybrid Nanoflower-Based Carbonic Anhydrase for Enhanced Biocatalytic Activity and Stability. ACS Omega 2018, 3, 18234–18241. [Google Scholar] [CrossRef] [Green Version]
  107. Heuer, J.; Kraus, Y.; Vučak, M.; Zeng, A.P. Enhanced sequestration of carbon dioxide into calcium carbonate using pressure and a carbonic anhydrase from alkaliphilic Coleofasciculus chthonoplastes. Eng. Life Sci. 2022, 22, 178–191. [Google Scholar] [CrossRef]
  108. Kondaveeti, S.; Abu-Reesh, I.M.; Mohanakrishna, G.; Bulut, M.; Pant, D. Advanced Routes of Biological and Bio-Electrocatalytic Carbon Dioxide (CO2) Mitigation toward Carbon Neutrality. Front. Energy Res. 2020, 8, 94. [Google Scholar] [CrossRef]
  109. Kim, H.S.; Hong, S.-G.; Yang, J.; Ju, Y.; Ok, J.; Kwon, S.-J.; Yeon, K.-M.; Dordick, J.S.; Kim, J. 3D-Printed interfacial devices for biocatalytic CO2 conversion at gas-liquid interface. J. CO2 Util. 2020, 38, 291–298. [Google Scholar] [CrossRef]
  110. Gao, H. A study on the spatio-temporal evolution of Tibetan inbound tourism flow network based on social networks. Proceedings of 2021 the International Conference on Environmental Remote Sensing and Big Data (ERSBD 2021), Wuhan, China, 29–31 October 2021; He, Y., Weng, C.-H., Eds.; SPIE: Bellingham, WA, USA, 2021; p. 42. [Google Scholar] [CrossRef]
  111. Iijima, T.; Yamaguchi, T. Efficient regioselective carboxylation of phenol to salicylic acid with supercritical CO2 in the presence of aluminium bromide. J. Mol. Catal. A Chem. 2008, 295, 52–56. [Google Scholar] [CrossRef]
  112. Iijima, T.; Yamaguchi, T. K2CO3-catalyzed direct synthesis of salicylic acid from phenol and supercritical CO2. Appl. Catal. A Gen. 2008, 345, 12–17. [Google Scholar] [CrossRef]
  113. Longtai, L.; Biao, G.; Xuebin, L.; Gang, W.; Limin, G. Research progress on hydrogenation of carbon dioxide. Ind. Catal. 2021, 29, 1–10. [Google Scholar] [CrossRef]
  114. Martins, J.A.; Miguel, C.V.; Rodrigues, A.E.; Madeira, L.M. Novel Adsorption-Reaction Process for Biomethane Purification/Production and Renewable Energy Storage. ACS Sustain. Chem. Eng. 2022, 10, 7833–7851. [Google Scholar] [CrossRef] [PubMed]
  115. Fu, L.; Ren, Z.; Si, W.; Ma, Q.; Huang, W.; Liao, K.; Huang, Z.; Wang, Y.; Li, J.; Xu, P. Research progress on CO2 capture and utilization technology. J. CO2 Util. 2022, 66, 102260. [Google Scholar] [CrossRef]
  116. Li, K.; Chen, J.G. CO2 Hydrogenation to Methanol over ZrO2-Containing Catalysts: Insights into ZrO2 Induced Synergy. ACS Catal. 2019, 9, 7840–7861. [Google Scholar] [CrossRef]
  117. Lau, N.-S.; Matsui, M.; Abdullah, A.A.-A. Cyanobacteria: Photoautotrophic Microbial Factories for the Sustainable Synthesis of Industrial Products. BioMed Res. Int. 2015, 2015, 754934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Dexter, J.; Fu, P. Metabolic engineering of cyanobacteria for ethanol production. Energy Environ. Sci. 2009, 2, 857. [Google Scholar] [CrossRef]
  119. Moret, S.; Dyson, P.J.; Laurenczy, G. Direct synthesis of formic acid from carbon dioxide by hydrogenation in acidic media. Nat. Commun. 2014, 5, 4017. [Google Scholar] [CrossRef] [Green Version]
  120. Zhao, G.; Joó, F. Free formic acid by hydrogenation of carbon dioxide in sodium formate solutions. Catal. Commun. 2011, 14, 74–76. [Google Scholar] [CrossRef]
  121. Chai, M.; Bazaz, S.R.; Daiyan, R.; Razmjou, A.; Warkiani, M.E.; Amal, R.; Chen, V. Biocatalytic micromixer coated with enzyme-MOF thin film for CO2 conversion to formic acid. Chem. Eng. J. 2021, 426, 130856. [Google Scholar] [CrossRef]
  122. Catizzone, E.; Bonura, G.; Migliori, M.; Frusteri, F.; Giordano, G. CO2 Recycling to Dimethyl Ether: State-of-the-Art and Perspectives. Molecules 2017, 23, 31. [Google Scholar] [CrossRef] [Green Version]
  123. Sun, K.; Lu, W.; Wang, M.; Xu, X. Low-temperature synthesis of DME from CO2/H2 over Pd-modified CuO–ZnO–Al2O3–ZrO2/HZSM-5 catalysts. Catal. Commun. 2004, 5, 367–370. [Google Scholar] [CrossRef]
  124. Fang, X.; Jia, H.; Zhang, B.; Li, Y.; Wang, Y.; Song, Y.; Du, T.; Liu, L. A novel in situ grown Cu-ZnO-ZrO2/HZSM-5 hybrid catalyst for CO2 hydrogenation to liquid fuels of methanol and DME. J. Environ. Chem. Eng. 2021, 9, 105299. [Google Scholar] [CrossRef]
  125. Ren, S.; Fan, X.; Shang, Z.; Shoemaker, W.R.; Ma, L.; Wu, T.; Li, S.; Klinghoffer, N.B.; Yu, M.; Liang, X. Enhanced catalytic performance of Zr modified CuO/ZnO/Al2O3 catalyst for methanol and DME synthesis via CO2 hydrogenation. J. CO2 Util. 2020, 36, 82–95. [Google Scholar] [CrossRef]
  126. Álvarez, A.; Bansode, A.; Urakawa, A.; Bavykina, A.V.; Wezendonk, T.A.; Makkee, M.; Gascon, J.; Kapteijn, F. Challenges in the Greener Production of Formates/Formic Acid, Methanol, and DME by Heterogeneously Catalyzed CO2 Hydrogenation Processes. Chem. Rev. 2017, 117, 9804–9838. [Google Scholar] [CrossRef]
  127. Mota, N.; Ordoñez, E.M.; Pawelec, B.; Fierro, J.L.G.; Navarro, R.M. Direct Synthesis of Dimethyl Ether from CO2: Recent Advances in Bifunctional/Hybrid Catalytic Systems. Catalysts 2021, 11, 411. [Google Scholar] [CrossRef]
  128. Pawelec, B.; Guil-López, R.; Mota, N.; Fierro, J.; Yerga, R.N. Catalysts for the Conversion of CO2 to Low Molecular Weight Olefins—A Review. Materials 2021, 14, 6952. [Google Scholar] [CrossRef] [PubMed]
  129. Ma, Z.; Porosoff, M.D. Development of Tandem Catalysts for CO2 Hydrogenation to Olefins. ACS Catal. 2019, 9, 2639–2656. [Google Scholar] [CrossRef]
  130. Guo, L.; Sun, J.; Ji, X.; Wei, J.; Wen, Z.; Yao, R.; Xu, H.; Ge, Q. Directly converting carbon dioxide to linear α-olefins on bio-promoted catalysts. Commun. Chem. 2018, 1, 11. [Google Scholar] [CrossRef] [Green Version]
  131. Qin, Y.; Wang, X.; Wang, F. Synthesis and properties of carbon dioxide based copolymers. Sci. Sin. Chim. 2018, 48, 883–893. [Google Scholar] [CrossRef] [Green Version]
  132. Liu, G.-L.; Wu, H.-W.; Lin, Z.-I.; Liao, M.-G.; Su, Y.-C.; Chen, C.-K.; Ko, B.-T. Synthesis of functional CO2-based polycarbonates via dinuclear nickel nitrophenolate-based catalysis for degradable surfactant and drug-loaded nanoparticle applications. Polym. Chem. 2021, 12, 1244–1259. [Google Scholar] [CrossRef]
  133. Shuai, L.; Xingyuan, M. Research Progress of Blocked Solvent-Free Polyurethane. Mater. Sustain. Dev. Green Manuf. Process. Mater. 2019, 33, 3892–3899. [Google Scholar]
  134. Ying, Z.; Zhang, C.; Jiang, S.; Wu, Q.; Zhang, B.; Yu, Y.; Lan, M.; Cheng, H.; Zhao, F. Synthesis of a novel hydrophobic polyurea gel from CO2 and amino-modified polysiloxane. J. CO2 Util. 2016, 15, 131–135. [Google Scholar] [CrossRef]
  135. Jiang, S.; Cheng, H.-Y.; Shi, R.-H.; Wu, P.-X.; Lin, W.-W.; Zhang, C.; Arai, M.; Zhao, F.-Y. Direct Synthesis of Polyurea Thermoplastics from CO2 and Diamines. ACS Appl. Mater. Interfaces 2019, 11, 47413–47421. [Google Scholar] [CrossRef] [PubMed]
  136. Zhao, L.; Sha, F.; Li, Y.; Zhang, J. Synthesis of polyurethane based on indirect utilization strategy of CO2. J. Environ. Chem. Eng. 2021, 9, 105309. [Google Scholar] [CrossRef]
  137. Spinner, N.S.; Vega, J.A.; Mustain, W.E. Recent progress in the electrochemical conversion and utilization of CO2. Catal. Sci. Technol. 2012, 2, 19–28. [Google Scholar] [CrossRef]
  138. Hong, S.G.; Jeon, H.; Kim, H.S.; Jun, S.H.; Jin, E.; Kim, J. One-pot enzymatic conversion of carbon dioxide and utilization for improved microbial growth. Environ. Sci. Technol. 2015, 49, 4466–4472. [Google Scholar] [CrossRef] [PubMed]
Figure 3. Schematic illustration of the substrate channeling effect between (A) two free-floating enzymes (E1 and E2) and (B) two proximate enzymes (E1/E2). Diffusion of the substrates (S1 and S2) and enzymes (E1 and E2) is assumed to be directional. Reproduced with permission from Hwang and Lee [71]. Copyright 2019 American Chemical Society.
Figure 3. Schematic illustration of the substrate channeling effect between (A) two free-floating enzymes (E1 and E2) and (B) two proximate enzymes (E1/E2). Diffusion of the substrates (S1 and S2) and enzymes (E1 and E2) is assumed to be directional. Reproduced with permission from Hwang and Lee [71]. Copyright 2019 American Chemical Society.
Catalysts 13 00611 g003
Figure 5. (A) Themultienzyme routefor thesynthesis of isobutyraldehyde from CO2. AlsS: acetolactate synthase; IlvC: acetohydroxy acid isomeroreductase; IlvD: dihydroxy-acid dehydratase; Kdc: 2-ketoacid decarboxylase, Reproduced with permission from Tong et al. [78] Copyright 2011 Wiley. (B) The multienzyme route for the synthesis of L-lactic acid from CO2 and ethanol. Originally published in Atsumi et al. [79] under Creative Commons Attribution 4.0 International License.
Figure 5. (A) Themultienzyme routefor thesynthesis of isobutyraldehyde from CO2. AlsS: acetolactate synthase; IlvC: acetohydroxy acid isomeroreductase; IlvD: dihydroxy-acid dehydratase; Kdc: 2-ketoacid decarboxylase, Reproduced with permission from Tong et al. [78] Copyright 2011 Wiley. (B) The multienzyme route for the synthesis of L-lactic acid from CO2 and ethanol. Originally published in Atsumi et al. [79] under Creative Commons Attribution 4.0 International License.
Catalysts 13 00611 g005
Figure 7. Schematic representation of CO2 capture, storage and utilization process. (A) Represents the CO2 sequestration process that involves the removal of CO2 from the flue gas that is in the absorber column and then desorbed in a heated stripper column to attain relatively pure CO2. (B) Depicts the process of bicarbonate formation. The immobilized CA (scrubber column) absorbed the released CO2, and the precipitation column’s carbonate ions were transformed into the compound CaCO3 in the presence of a CaCl2 solution. Reproduced with permission from Kanth et al. [97] Copyright 2013 Wiley.
Figure 7. Schematic representation of CO2 capture, storage and utilization process. (A) Represents the CO2 sequestration process that involves the removal of CO2 from the flue gas that is in the absorber column and then desorbed in a heated stripper column to attain relatively pure CO2. (B) Depicts the process of bicarbonate formation. The immobilized CA (scrubber column) absorbed the released CO2, and the precipitation column’s carbonate ions were transformed into the compound CaCO3 in the presence of a CaCl2 solution. Reproduced with permission from Kanth et al. [97] Copyright 2013 Wiley.
Catalysts 13 00611 g007
Figure 8. Schematic diagram illustrating the three major pathways for producing olefins from CO2: (1) modified Fischer-Tropsch to olefins (FTO) route; (2) methanol to olefins (MTO) route; and (3) direct olefin formation over multifunctional, promoted catalysts. RWGS—reverse water-gas shift. Adapted from [128].
Figure 8. Schematic diagram illustrating the three major pathways for producing olefins from CO2: (1) modified Fischer-Tropsch to olefins (FTO) route; (2) methanol to olefins (MTO) route; and (3) direct olefin formation over multifunctional, promoted catalysts. RWGS—reverse water-gas shift. Adapted from [128].
Catalysts 13 00611 g008
Figure 9. Chemo-enzymatic CO2 utilization and conversion [27]. Reproduced with permission from Elsevier 2021.
Figure 9. Chemo-enzymatic CO2 utilization and conversion [27]. Reproduced with permission from Elsevier 2021.
Catalysts 13 00611 g009
Table 2. List of products formed via chemo-enzymatic approach of CO2 conversion.
Table 2. List of products formed via chemo-enzymatic approach of CO2 conversion.
CatalystProductsRef.
Carboxydothermus hydrogenoformans CO Dehydrogenase I [Ch Ni-CODH I]Carbon monoxide (CO)[24]
Dehydrogenases in an ALG−SiO2Methanol[83]
Formate DehydrogenaseFormate[41]
NitrogenaseMethane[84]
Carbonic anhydraseBicarbonate (HCO3)[85]
Nanocatalyst (Na–Fe3O4/HZSM-5)Gasoline[86]
Nanoparticles (PdZn/TiO2)Dimethyl Ether[87]
Inorganic catalyst ZnO-ZrOStarch[88]
Formolase(FLS) and D-fructose-6-phosphate aldolase (FSA)Carbohydrate (l-Erythrulose)[89]
Hybrid nanocatalyst (C@Fe–Al2O3)Salicylic acid[90]
PdCu alloy nanoparticlesUrea[91]
Metallic heterogenous catalystCyclo carbonates[92]
Cu–Ru–Metallic Organic Framework (MOF)Ethanol[92]
Platinum/cobalt catalystsButanol-based fuel[93]
Table 3. Advantages and disadvantages of various CO2 conversion approaches.
Table 3. Advantages and disadvantages of various CO2 conversion approaches.
ApproachesAdvantagesDisadvantagesRef.
Electrochemical
  • Recyclable electrolytes
  • High selectivity
  • Long term stable operation
  • Easy to scale-up
  • Low catalyst lifespan
  • Economically not feasible
  • Require more electrical energy
[94]
Thermal
  • High yield production
  • Require high temperature
  • Low catalyst stability
[34,95]
Chemo-enzymatic
  • Bulk chemicals production
  • High selectivity and yield
  • Reaction requires less variable conditions
  • Require high temperature
[27,85]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Markandan, K.; Sankaran, R.; Wei Tiong, Y.; Siddiqui, H.; Khalid, M.; Malik, S.; Rustagi, S. A Review on the Progress in Chemo-Enzymatic Processes for CO2 Conversion and Upcycling. Catalysts 2023, 13, 611. https://doi.org/10.3390/catal13030611

AMA Style

Markandan K, Sankaran R, Wei Tiong Y, Siddiqui H, Khalid M, Malik S, Rustagi S. A Review on the Progress in Chemo-Enzymatic Processes for CO2 Conversion and Upcycling. Catalysts. 2023; 13(3):611. https://doi.org/10.3390/catal13030611

Chicago/Turabian Style

Markandan, Kalaimani, Revathy Sankaran, Yong Wei Tiong, Humaira Siddiqui, Mohammad Khalid, Sumira Malik, and Sarvesh Rustagi. 2023. "A Review on the Progress in Chemo-Enzymatic Processes for CO2 Conversion and Upcycling" Catalysts 13, no. 3: 611. https://doi.org/10.3390/catal13030611

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