Sabiha Sultana
,
Prakash Chandra Sahoo
*,
Satyabadi Martha
and
Kulamani Parida
*
Centre for Nano Science and Nano Technology, ITER, Siksha ‘O’ Anusandhan University, Bhubaneswar – 751030, Odisha, India. E-mail: kulamaniparida@soauniversity.ac.in; prakash200920@gmail.com; Fax: +91-674-2351217; Tel: +91-674-2350181
First published on 18th April 2016
Rapid climate change and increase in the demand for hydrocarbon fuels have made the research community to recheck the carbon dioxide (CO2) concentration and its utilization in a green and efficient way. One of the best ways to utilize CO2 so far considered is its conversion back into fuels using various strategies such as thermal, photochemical, electrochemical and enzymatic. In this critical review the concepts and mechanism of CO2 reducing biocatalysts have been summarized. Specifically, more emphasis was given to enzymes related to the selective conversion of CO2 to fuels (popularly known as artificial photosynthesis). Emphasis has been given to single enzyme and multi-enzymatic systems for CO2 conversion to carbon monoxide, methanol, and formic acid/formate. Semiconducting nano-material integrated enzymatic systems and their mechanism are reviewed and described in detail. Finally, the ultimate challenges underlying the design of next generation enzymatic catalysts are explored.
Carbon dioxide, in general, is a very stable and non-flammable compound. At normal pressure and temperature, it is odourless, colourless, and heavier than air, with a slightly pungent odour. It has wide applications and it is used for many purposes in many industries, such as in cooling (solid ice), agro-production for conservation, and for medical applications. It is known to be produced during fermentation, combustion and putrefaction. Below its immediately lethal concentration, CO2 has long been considered as a neutral compound in the human body. The maximum acceptable concentration of CO2 lies between 0.5 and 3% but long term exposure of a minimum concentration of CO2 may be dangerous for health.10 So a rise in CO2 concentration in the atmosphere is a warning for the future and this was debated extensively at the Paris summit in 2015 (COP-21). The remaining anthropogenic CO2 should be minimized in order to avoid any disturbances in natural cycles, the environment and also to human and animal life.
In order to alleviate anthropogenic CO2 emission, different strategies have been proposed by various scientific communities. Among them the first way proposed is to reduce the direct emission of CO2, which is quite impossible because of insufficient alternative energy sources. However, because of the rising population, plentiful energy sources are required and their demand is only fulfilled by using fossil fuels11 (Fig. 1b) which ultimately leads to a huge anthropogenic CO2.
Carbon dioxide capture and storage is a second useful technique for CO2 minimization. At present, the readily available technology to tackle CO2 associated global warming problem is the carbon capture and storage (CCS) process. In this technology CO2 is captured and stored in risk free, secure places such as depleted coal seams, under the sea at deep levels and so on.12–14 It is limited because of the high energy requirement for gas transportation and compression and there is also a risk of leakage to environment. The continuous catalyst regeneration (CCR) process (i.e., capturing the CO2 and reusing it) has been considered to be a valuable and economical choice in comparison to the traditional CCS process. Besides economic benefits, the socio-political benefits also come in terms of a positive image for companies adopting policies of reusing the CO2 generated from fossil fuels.9
A third alternative is to convert the anthropogenic CO2 in to useful chemicals by reducing it. Various methods have been proposed for this: electrochemical conversion,15–18 photochemical conversion,19–22 thermochemical23–25 and so on. However, these technologies have drawbacks such as a high temperature and electricity requirements, drastic conditions, low selectivity of products (formation of more than one product unselectively) and so on.
Fossil fuels are the most important source of energy and human dependence on them increases day by day because of their high energy density, high stability and because they are readily available and easy to store and transport.26 But these sources are non-renewable and their amount is depleted continuously. So nowadays the search for inexpensive and plentiful energy supplies appears to be at odds with climate change mitigation commitments.27 So, different fuels have been proposed as alternatives to fossil fuels. Hydrogen (H2) has emerged as an appealing energy resource for the future generation energy systems. It has the capability to minimize pollution as well as rechecking the greenhouse gas emissions, that since the beginning of the industrial revolution have been plaguing mankind.28 But there are some disadvantages for H2 being used as a fuel. Its production is a highly labour intensive, energy consuming process. It is neither acceptable as a means to save energy nor for its subsequent use as a fuel. The handling of potentially explosive and volatile (boiling point: −253 °C) H2 gas which has negative Joule–Thomson coefficient, requires special conditions such as high pressure, use of special materials to minimize leakage and diffusion, and extensive safety precautions. Even so, its potential use as a fuel is limited because any leaks could be potential explosion hazards,29 whereas hydrocarbon fuels do not suffer from these severe drawbacks. For a long-term solution to mankind's need of fuels as well as for efficient storage of energy generated, a non-fossil fuel source is needed. The utilization and recycling of the anthropogenic atmospheric carbon CO2 itself through its reduction to use of hydrocarbon fuel such as methanol (CH3OH), formic acid (HCOOH), carbon monoxide, formaldehyde (HCOH) offers a new feasible alternative.29
Very recently CO2 utilisation by a process of reduction has emerged as an exciting field of research. The use of CO2 by green plants for photosynthesis is an excellent method for its utilisation, whereas the natural processes are very slow and complicated. A natural enzymatic system has been established as an efficient green method which selectively generates products with high yield at milder conditions. Enzymatic reduction offers a green, promising approach for CO2 reduction. In general, an enzymatic reduction system is composed of several enzymes individually or combined with a number of other enzymes, in order to perform their action on a substrate. In a CO2 reduction system enzymes are sometimes used as an electrocatalyst or they are integrated with a photocatalyst which harvests sunlight and convert it into chemical energy in the form of hydrocarbon fuel. This photocatalyst integrated system offers an artificial photosynthetic route which will become a key system for future generation of solar fuels. In this critical review the recent trends of the enzymatic CO2 reducing system have been summarized and are discussed with a detailed and mechanistic approach.
Reduction of CO2 is a highly unfavourable endergonic process. The majority of CO2 reduction processes (eqn (2)–(6)) is highlighted between −0.24 V and −0.6 V, but the major problem is the single electron reduction (eqn (1)) of CO2 at −1.90 V which makes the single electron reduction reaction extremely unfavourable. In addition, there is a huge kinetic overvoltage required for the one electron reduction because of the structural differences between bent CO2˙− and linear CO2 which can be clearly seen by Walsh's diagram (Fig. 2a).30
Reactions, E° (V vs. NHE) |
CO2 + 2e− → CO2˙−, −1.90 | (1) |
CO2 + 2H+ + 2e− → HCOOH, −0.61 | (2) |
CO2 + 2H+ + 2e− → CO + H2O, −0.53 | (3) |
CO2 + 4H+ + 4e− → HCHO + H2O, −0.48 | (4) |
CO2 + 6H+ + 6e− → CH3OH + H2O, −0.38 | (5) |
CO2 + 8H+ + 8e− → CH4 + 2H2O, −0.24 | (6) |
![]() | ||
Fig. 2 (a) Walsh diagram for the total energies upon CO2 bending and (b) molecular orbitals of CO2 at certain OCO angles (reprinted with permission from ref. 30, license no. 3845730801023). |
A large amount of energy input is required to convert CO2 to other reduced species. But the energy involved in the formation of CO2˙− is much greater than that for the other reduced species. It is may be because of to the sizable reorganization energy attendant with the conversion of the linear CO2 molecule to the bent radical anion CO2˙−. So its reduction is typically identified by a considerable decrease in the OCO angle besides the expected C–O bond lengthening. Originally for CO2 lowest unoccupied 2πμ molecular orbital is doubly degenerate and is composed of two antibonding π-orbitals which are orthogonal in nature. The injection of a single electron to the lowest unoccupied molecular orbital (LUMO) is energetically difficult, so a large over voltage is required. However, after the injection of electron, CO2 bending (134°) results in lowering the energy of its LUMO (2πμ), which is originally placed high enough in linear form (Walsh's diagram) as shown in Fig. 2b. Thus the enhancement in the electron accepting tendency of CO2 occurs which is favourable for other reduction. Such geometric distortion in fact entails a significant energetic expense as evidenced by the sufficiently negative potential for the one electron reduction.30,32–34
The reduction reactions of CO2 which are listed in eqn (2)–(6) all take place at much lower potentials than the first single electron reduction reactions. This may be because each of these reduction reactions involves the transfer of multiple reducing equivalents, and as a result, the formation of the high energy intermediate, CO2˙− will be circumvented. Additionally, each of these reduction reactions are associated with the transfer of an equal number of protons. More stable intermediates are formed for these multi-electron proton coupled electron transfer reactions (PCET) than the single electron reduction of CO2. Four of the last PCET reactions (eqn (3)–(6)) result in the formation of reduced carbon containing species in addition to the formation of at least one equivalent of water (H2O). Because water is a very stable molecule, its formation helps to offset the energy needed to drive these CO2 reduction processes.34 This explains the kinetics of CO2 reduction reactions, but the thermodynamics behind these reactions are quite different.
Because of the high stability (ΔG°), a substantial energy input, optimized reaction conditions, and catalysts with high stability and activity are required to convert CO2 into value added chemicals.8 According to the Gibbs–Helmholtz relationship for chemical reactions, differences between the Gibbs' free energy of reactants and products under certain conditions mainly drives the chemical reactions. The terms associated with Gibbs free energy (TΔS° and ΔH°) are not favourable for converting CO2 into other valuable products.9 A high energy input is required for dissociating the bonds in CO2 as the carbon–oxygen bonds are very strong. Similarly the entropy term (TΔS°) is also not a thermodynamic driving force for any of the reactions involving CO2. In fact the enthalpy term (ΔH°) can be conveniently considered for assessing the thermodynamic feasibility and stability of any CO2 conversion. However, it was suggested that a positive change in free energy should not be considered as sufficient reason for not pursuing potentially useful reactions of CO2 reduction which would give value added products. Even though ΔG° gives discouraging values for the formation of products at equilibrium through the relationship (ΔG° = −RTln
k), kinetics do favour certain reactions. Thus, in order to convert CO2 into value added chemicals, a good catalytic system is required. The thermodynamics of CO2 and its complications can easily be understood by studying the Frost diagram (Fig. 3). According to the Frost diagram, on the vertical axis, elements of the most stable oxidation state are placed. Oxidants and high energy fuels are placed higher up at the right and left sides of the diagram, respectively. The slope of the line joining two couples gives the reduction potential of their inter-conversion. The intermediates present above these gradients are quite unstable with respect to disproportion. Those with steeper slopes require more energy input for their interconversion. Also, the Frost diagram shows the concept of overpotential. It is the angular difference between the line joining of a couple and the potential, i.e., that which is actually required to drive the reaction. This is generally wasted energy and it can practically be influenced by the catalyst which will stabilise the inbetween intermediate. From Fig. 3 it is seen that CO2 reduction is much more difficult than water splitting. It has been seen that a single electron reduction of CO2 is an uphill process, as the large, albeit minimal, overpotential is required for continuing the reaction. Whereas, reduction to formate and CO is much more favourable as it has gone through the PCET reactions. However, reduction of CO2 to HCOH is the most energetically costly step, but the later reduction steps serve mainly to store equivalents of H2 generated upon conversion from CO2 to CH3OH and methane. In this review, how the enzymes (which are substrate specific in nature) facilitate these PCET reactions by avoiding the unwanted intermediates efficiently will be discussed.35–37
![]() | ||
Fig. 3 Frost diagram showing redox reactions of CO2, H2O, H2 and their intermediate products (reprinted with permission from ref. 37). |
CO2 + 2H+ + 2e− ↔ CO + H2O, CODH, electron donor |
Two types of CODHs have so far been identified in nature by the distinct structure of their active site. Some aerobes use enzymes containing a μ-sulfido bridged copper–sulfur–molybdenum (Cu–S–Mo) centre.48 However, some anaerobic micro-organisms contain nickel (Ni) and iron (Fe) as shown in Fig. 4, which catalyses the reversible redox reaction of CO/CO2 and the synthesis of acetyl-coenzyme A (acetyl-CoA).49 CODHs have a α2β2 quaternary structure with metal ions organized into four types of cluster: A, B, C, D. The C-cluster is only responsible for the CO/CO2 chemistry.50 The active site [NiFe] CODHs (C-cluster) contains a co-ordinatively unsaturated Ni centre bound to three sulfur ligands and a dangling Fe centre which is co-ordinated by cysteine (Cys), histidine (His), sulfido and an H2O/OH− ligand. Both ions are bridged by an iron(II, III) sulfide (Fe3S4) cluster and are positioned in close proximity. At first, CO2 is bound to the Ni centre and stabilized by the Fe centre, only then does the reduction occur.48,50 In terms of the subunit components and metabolic role of the enzymes, Ni-CODHs have been divided into four classes. Among them class IV and class III enzymes are involved in the process of reduction where all the class IV enzymes contain a dimeric structure, each monomer consists of a C-cluster [Ni3Fe4S] and a dangling Fe or extra cuboidal pendant. The Ni-CODH family is extended to include the class III enzymes known as CODH/acetyl-COA synthetase (ACS) (EC 1.2.7.4) a bifunctional 310 kDa α2β2 tetramer in which an ACS is strongly associated with Ni–Fe–CODH.52 The CODH/ACS comprises four subunits: two CODH proteins and two ACS proteins. The β subunits of ACS/CODH are homo-dimeric in nature which is homologous to simple CODH enzymes found in carboxydotrophic bacteria. Each β subunit contains a novel active site C-cluster [Ni–Fe4–S4-or-5] and a [Fe4–S4] B-cluster. The two β subunits are bridged through a single [Fe4–S4] D-cluster and each α subunit of ACS/CODH contains the A-cluster, a single metal center, which is responsible for the ACS activity.30,50–53 Generally CODH from Moorella thermoacetica50,52 and Carboxydothermus hydrogenoformans (Ch)51,53–57 shows CO2 reductase activity.
![]() | ||
Fig. 4 Structure of CODH showing its catalytic active site (reprinted with permission from ref. 30, license no. 3845730801023). |
CO2 + NADH ↔ HCOOH + NAD+, FDH |
The diversity of reactions as well as the enzyme composition depends upon electron acceptors such as quinine, ferredoxin, NAD(P)H and 5-deazaflavin. Brief studies on these enzymes have contributed to the genetic and biochemical understanding of Fe, Mo, selenium (Se) and tungsten (W), in biology. FDHs are active over a wide range of pH and can therefore be coupled with almost any other dehydrogenases. FDHs are a suitable biocatalyst for the efficient synthesis of bulk drug substances, agricultural products, chiral intermediates and fine chemicals.48,58,59 In general, FDHs are categorised in two parts according to the presence of a metal ion in their catalytic site.
![]() | ||
Fig. 5 (A) Topological diagram for CbFDH. (B) A monomer, and (C) the dimer demonstrating the tight interaction between the NAD+ binding domains of the subunits (reprinted with permission from ref. 48, license no. 3845740593135). |
In the active site of Mo/W-FDH, one Mo/W atom is coordinated by the cis-dithiolene group of one or two pyranoprotein cofactor (only two PGD for W) molecules and these are mainly responsible for binding CO2. The coordination sphere is completed with an O, S or Se atom in a diversity of arrangements that determines the classification of molybdenum-enzymes into three big families xanthine oxidase, sulfite oxidase and dimethylsulfoxide reductase (DMSOR). In W-FDH (EC 1.2.1.43) the coordination sphere is completed only with O and/or S atoms from terminal groups such as amino acid side chains in the same trigonal prismatic geometry found in DMSOR family members. The electrons are necessary to carry out the reaction involving these enzymes and are transferred from the electron donor or to the electron acceptor, through different redox centres, such as the Fe–S centre, haem and flavins.72
Generally the NAD+–Mo–FDH (EC 1.1.99.33) from Cupriavidus oxalaticus, Rhodobacter capsulatus,73 Acetobacterium woodii,74 and Escherichia coli75 are involved in CO2 reductions. These are multimeric containing several flavins and Fe/S centres. R. capsulatus FDH especially, is a heterotrimer (αβγ)2 (345 kDa) containing an Mo centre, four [4Fe–4S], one [2Fe–2S] centres and one flavin mononucleotide. The Mo centre is generally responsible for formate conversion. Similarly NAD+–W–FDH from Syntrophobacter fumarodoxin,76,77 M. thermoacetica78 and Clostridium carboxidivorans79 is involved in the CO2 reduction process. M. thermoacetica is a dimer of a heterodimer (αβ)4 (95 and 75 kDa) containing a W centre with a cysteine ligand and several Fe/S centres.
HCOOH/HCOO− + NADH ↔ HCOH + NAD+, FaldDH |
FaldDH has been found in a wide variety of organisms such as rats, humans, bovines, C. boidinni, and Pseudomonas putida. The latter organism was originally numbered as EC 1.2.1.1 but with its revised nomenclature it has placed in EC 1.2.1.46.81 It is categorized under the zinc containing medium chain ADH family. Almost all FaldDHs require NAD+ and/or glutathione to activate the reaction. But the FaldDH extracted from P. putida is a structurally distinguished enzyme that can operate the reaction without the external addition of glutathione80 and it can be used as an intermediate enzyme in the CO2 reduction process.82–84
FaldDH from P. putida is an excellent enzyme used in catalytic systems. Its molecular mass is nearly equal to 42 kDa. In the active form, it possesses a homo tetramer of identical subunits each of which contains two zinc ions and 398 amino acid residues. Each subunit comprises two firmly bound zinc ion and two domains which are separated by a cleft containing a deep pocket, and one of the catalytic zinc ions is found at the bottom of the cleft. The catalytic domain and cofactor binding domains are unequal in shape and size as shown in the (Fig. 6). The larger catalytic domain consists of 231 amino acid residues whereas the smaller cofactor binding domain contains only 167 residues out of the 398 residues. The secondary structure of the cofactor binding domain comprises a six stranded β-sheet which is arranged in parallel in the centre of domain. The six stranded β-sheets are surrounded by 4 α-helices which shows a Rossmann fold comprised of a (βαβαβ)2 unit.80,81
![]() | ||
Fig. 6 Ribbon view of the NAD+ complex of the P. putida FaldDH enzyme subunit (α-helices (violet) and β-strands (yellow)) (reprinted with permission from ref. 81, license no. 3845740809509). |
HCOH + NADH ↔ CH3OH + NAD+, ADH |
Four different families of ADH were identified in plants, eukaryotic micro organisms, vertebrates and prokaryotic bacteria.85 It is responsible for the breakdown of toxic alcohols to aldehydes in humans and animals. Yeast ADH (YADH) shown in Fig. 7 is included in the third family and was one of the first enzymes which were isolated and purified.86 It comprises a group of isoenzymes that catalyse the oxidation of alcohols to aldehydes in the presence of suitable cofactor/electron donor. The three isoenzymes (I, II, III) of ADH are found in, S. cerevisiae (baker's yeast). YADH-1 is generally expressed during anaerobic fermentation85 whereas YADH-2 is another cytoplasmic form which is repressed by glucose and YADH-3 is found in the mitochondria.87 YADH-1 plays a vital role in the activity of growing S. cerevisiae. YADH-1 and YADH-2 have very similar characteristics, whereas YADH-2 is more substrate specific in nature.88 YADH-1 is a tetramer, consisting of four identical subunits: each subunit consists of a single polypeptide chain with 347 amino acids, with a molecular mass of 36 kDa. The subunits of the yeast enzyme are divided into two domains such as the catalytic domain and the coenzyme-binding domain. One domain binds the coenzyme and the other provides a ligand to the catalytic zinc, as well as to most of the groups that control substrate specificity.88 In most of the CO2 reducing multiple enzymatic systems, YADH-1 is used.
Both FaldDH and ADH cannot act upon CO2 individually; these two enzymes combined with FaldDH give CH3OH from CO2 which is spatially categorised under the multiple enzymatic system. So in this review an outline of an enzymatic system is given, which gives a clear description of the enzymatic system for CO2 reduction (Scheme 1).
The first report of CO2 reduction to CO with an appropriate mechanism on CODH was suggested by Anderson and Lindahl.90 They stated that only two reduced states of the C-cluster in CODH are capable of CO2 reduction. According to their proposed mechanism, the CO2 molecule bound to the Cred1 state of the C-cluster as shown in Scheme 2. Cred2 was then formed by the reduction of the Ni centre by dithionite. According to Anderson and Lindahl, catalysis required one active site C-cluster, one electron transfer B-cluster and one unidentified redox agent X. CO2 was first bound and then was reduced by Cred2 then Cint was produced after a one electron transfer to either the species X or the B-cluster. Later Fraser and Lindahl91 stabilized the Cint species using the Ti3+/argon substrate and this supported the mechanism described by Anderson and Lindahl. Since then, much extensive study and research has been done by the various research groups.52,89–92
![]() | ||
Scheme 2 Model of CO oxidation/CO2 reduction catalysis using CODHs (reprinted with permission from ref. 90, Copyright American Chemical Society 1996). |
The correct mechanism is that, when CO2 is added to the reduced state of the C-cluster CODH, it first binds to the Ni centre. Because of the binding, minimal geometry changes occur and the fourth position around the Ni occupies a square planar geometry from planar T-shaped geometry. Thereafter, a water molecule is lost from Fe and this formed the Fe–O–C bond. Because of the coordination of CO2, iron acts as the Lewis acid whereas nickel acts as a Lewis base, whereas the resulting CO2 adduct is further stabilized through H2-bonding formed by the protonated histidine residue. The Fe3S4 framework holds Fe and Ni in place so that the position of these metals remain unchanged after or before CO2 addition and this cluster also acts as an electronic buffer which stabilizes the further electronic charges arising on Ni and Fe during the catalytic cycle. The formation of a Ni(II) CO species occurs with a corresponding cleavage of the C–O bond and loss of a water molecule. This Ni(II) CO species readily loses CO and enters the catalytic cycle by accepting an H2O molecule and completes the cycle as shown in Scheme 3.47,93,94
![]() | ||
Scheme 3 Proposed mechanism for the conversion of CO2 to CO using CODH (reprinted with permission from ref. 94. Copyright American Chemical Society 2013). |
![]() | ||
Fig. 8 (a) Diagrammatic representation of the surface of MOx (TiO2 and CdS) where an electron accumulation layer is formed, when Eappl is lower than Efb. (b) Cyclic voltammograms of unmodified and enzyme modified electrodes in 0.2 M 2-(N-morpholino)ethanesulfonic acid (MES), at pH – 6.0, 20 °C (reprinted with permission from ref. 55. Copyright American Chemical Society 2013 (open access)). |
![]() | ||
Fig. 9 (a) Schematic representation of CO2 photoreduction system using CODH attached to RuP modified TiO2 NPs. Here the active site is clearly shown (reprinted with permission from ref. 56. Copyright American Chemical Society 2010). (b) Representation of MOx modified with CODH1 and Ru-sensitizer. (c) Graphical representation of the activity of different MOx. Among them TiO2 (P-25, Degussa) exhibits highest activity (reprinted with permission from ref. 51). |
Woolerton et al.51 also studied the effect of different metal oxide nanoparticles (NPs) attached with CODH and a Ru photo-sensitizer. They reported that the core mechanism is the same as the one found in the previous study. Here, with the irradiation of visible light, the ruthenium phosphate (RuP) gets sensitized and injects electrons into the CB of the semiconductor photocatalyst shown in Fig. 9b. These high energetic electrons are then able to enter the enzyme through the D-cluster and reach the active site via the B-cluster, where CO2 is reduced to CO. It was found that P25 (TiO2) NPs modified enzyme shows the highest rate of CO production. From Fig. 9c it is shown that a rutile TiO2 enzyme-based system showed no detectable CO production, because of the low uptake of both enzyme and photosensitizer component by the TiO2 NPs. Another reason is that the rutile CB edge position which is at −0.32 versus SHE at pH 6 is too low for CO2 reduction. Anatase TiO2 NPs have shown similar results in P25 (TiO2) because the catalyst consists of 80% of the anatase form of TiO2. Similarly, ZnO NPs showed less production yield even though its CB position is nearer to the anatase TiO2. As ZnO has a higher isoelectronic point than P25 at pH 6, CODH adsorption will be favoured because of its surface positive charge. A small amount of detectable CO was formed which is 10% less than that of P25. The low solubility of ZnO in the buffer system leads to desorption of enzyme and photosensitizer resulting in decreased production yield. As shown in Fig. 9c, an SrTiO3-based enzyme system shows a poor production yield of CO. This is because of the small potential difference between the CB of the photocatalyst and the excited state of the dye, which could not drive a rapid injection of electrons into the semiconductor. It was observed that a change of sacrificial electron donor also affected the product yield. For the standard system P-25TiO2, the turnover frequency at 20 °C is 0.1 s−1. Chaudhary et al.57 used an alternative metal compound CdS as a photocatalyst which was then combined with the CODH enzyme without any photosensitizer, because the solvent, MES, acts as an electron donor which quenches the hole produced from the semiconductor. They found that the particle size of the photocatalyst has a remarkable effect on CO production. They used CdS quantum dots (QDs), CdS NPs and CdS quantum dot calcined particles and attached these with the enzyme and a higher rate of photoreduction of CO2 was observed using the CdS nanorod system (turnover frequency ∼ 1.23 s−1 per enzyme molecule) because of the lower extraction recombination probability.
![]() | ||
Scheme 4 Catalytic mechanism for the reduction of CO2 to formate using a metal independent NAD+-dependent FDH (reprinted with permission from ref. 95). |
The first direct CO2 reduction to formate was carried out by Ruschic et al. by using FDH and NADH, with a turnover frequency in the range of 3 s−1.62 Later Parkinson and Weaver63 intensively studied the reaction using p-type indium phosphide as a photocathode and they successfully used photogenerated electrons to reduce MV, which mediates the FDH linked reduction of CO2 to formate. The group reported an 80–93% product formation. Several studies have been done by different research groups using FDH obtained from C. boidinii.61 A paper by Choe et al.64 reported that at neutral pH, Thiobacillus FDH (TbsFDH) is more active than the conventionally used CbFDH for CO2 reduction. TbsFDH showed the highest CO2 reducing activities, i.e., 5.8-fold higher than those obtained with CbFDH. This phenomenon was explained by the structural differences and showed that TbsFDH has an elongated N- and C-terminal loop on its protein backbone. However, the turnover rate is still low for this type of mobilised system, so FDH was first immobilised on alginate silica (ALG–SiO2) hybrid gel. One of the advantages of this system is that these hybrid gels immobilised the enzyme so that the enzyme can be used multiple times in repeated catalytic cycles. The highest yield of HCOOH in this case was up to 96% and after 10 cycles the relative activity of the immobilized FDH can be maintained as high as 69%.65 Wang et al. reported that a combination of the two enzymes of FDH with carbonic anhydrase improved the product formation yield.66 They explained that CO2 might be transferred to more soluble HCO3− when carbonic anhydrase was added to the system resulting an increase in the hydration rate of CO2. So if there is a higher number of substrates available for the enzyme, the production rate of HCOOH increases by a factor of 4.2-fold compared to that obtained with a single FDH.
FDHs can also combine with a photocatalyst to improve their activity and product yield. Miyatani and Amao67 reported an enzymatic photochemical system containing FDH, MV2+, zinc tetrakis(4-methylpyridyl)porphyrin (ZnTMPyP) in the presence of triethanolamine (TEOA) as an electron donator. They described that upon irradiation of light on a solution containing TEOA, electrons were released, which was mediated by MV2+ and ZnTMPyP and then they moved to the FDH active site where CO2 was reduced to HCOOH. NADH is mainly used in a catalytic system as a terminal electron donor, which is highly expensive. However, in this system instead of natural NADH they used MV2+ and the product formation was found to be 62 μmol which is comparable to that obtained with the more expensive system. Amao et al.68 reported a system (artificial leaf) which consisted of chlorin e6, a photosensitized dye together with the electron mediator viologene and FDH on a silica gel substrate to obtain HCOOH from CO2. Upon irradiation of light the electron flows from the dye to MV2+ then to FDH in presence of NADPH where it reduces CO2 in a saturated bis–Tris buffer solution of pH 7 as shown in Fig. 10a. In a recent paper by same group,69 H2 and HCOOH production was reported by using a coupled system containing glutamate dehydrogenase (GDH) and FDH. Where, simultaneous glucose oxidation by GDH, H2 and HCOOH production with platinum (Pt) NPs and FDH was observed, as shown in Fig. 10c. Using the visible light induced photoreduction of MV2+ by photosensitization of zinc tetraphenylporphyrin tetrasulfonate (ZnTPPS), 0.2 μmol of HCOOH was produced.
![]() | ||
Fig. 10 (a) Schematic representation of the artificial leaf by mimicking photosystem II (PSII), producing HCOOH (reprinted with permission from ref. 68). (b) A graphene-based photocatalyst, CCGCMAQSP with Rh-photosensitizer catalyzed NADH regeneration as well as artificial photosynthesis of HCOOH from CO2 under visible light (reprinted with permission from ref. 70. Copyright American Chemical Society 2012). (c) Visible light induced H2 and HCOOH production with Pt NPs and FDH (reprinted with permission from ref. 69, license no. 3845750285042). |
In order to regenerate NADH in the same catalytic system, Yadav et al. reported on a photocatalytic-enzyme coupled system for HCOOH production.70 In this they used the photocatalyst, chemically converted graphene coupled multi-anthraquinone-substituted porphyrin (CCGCMAQSP) with NADH, TEOA as an electron donor, a photosensitizer rhodium (Rh) complex and FDH. The incident light is first absorbed by MAQSP, where the photoexcitation occurs and through the stable amide bond the photoexcited electrons move into the CCG, where the TEOA acts as a hole scavenger. The extremely high charge carrier mobility and huge surface area of graphene increase the probability for electrons to be transferred into the Rh complexes, eventually accelerating the chemical regeneration of NADH as shown in Fig. 10b. In another paper by Yadav et al.71 use of a robust photocatalyst called: chemically converted graphene 1-picolylamine-2-aminophenyl-3-oxy-phenyl-4,4,4′-difluoro-1,3,5,7-tetramethyl-2,6-diethyl-4-bora-3a,4a-diaza-S-indacene-triazine (CCG-BODIPY) was proposed. Using CCG-BODIPY as a photocatalyst with other enzymes they effectively regenerated NADH in the medium with a cumulative HCOOH production yield.
![]() | ||
Scheme 5 Catalytic cycle of CO2 reduction to formate ion via FDH (reprinted with permission from ref. 95). |
FDH obtained from the anaerobic syntrophic S. fumaroxidans shows CO2 reduction activity.76,77 De Bok et al.76 reported both CO2 reduction as well as formate oxidation by taking FDH1 and FDH2 from this anaerobe. They found that FDH1 shows a higher CO2 reduction reaction than formate oxidation, where as FDH2 exhibited a 30 times higher formate oxidation activity than CO2 reduction. Similarly, Reda et al.77 reported the use of FDH1 on a PGE electrode surface. This electrocatalyst showed efficient conversion of CO2 to formate with minimal over potential in milder conditions at a turnover rate of ≈0.5 × 103 s−1. Another FDH from C. carboxidivorans was demonstrated by Alissandratos et al.,79 which showed CO2 reductase activity at a kcat/km of 1.6 s−1 mM−1 and in this case the FDH is an NAD(P)H dependent one. They showed that the FDH of C. carboxidivorans shows that the CO2 reductase activity is more efficient than that of the C. boidinii FDH. A remarkable H2 dependent and specific CO2 reductase complex was developed by an acetogen called A. woodii which directly couples the dihydrogen oxidation with the CO2 reduction. This notable reductase complex allows the CO2 reduction by dihydrogen with a turnover frequency of 28.22 s−1.74 Bassegoda et al.75 reported a Mo-FDH extracted from E. coli (EcFDH-H), which was able to catalyse the reduction of CO2 efficiently and specifically. They found that compared to the FDH of S. fumaroxidans (SfFDH1) the EcFDH-H is strongly biased towards CO2 reduction at pH 7.5. Hartmann and Leimkuhler73 studied oxygen tolerant Mo FDH from R. capsulatus which was efficiently able to catalyse both CO2 reduction and formate oxidation in solution. However, the CO2 reduction reaction was observed at a turnover rate of 1.5 s−1, and was much slower than the observed (∼20 times slower) formate oxidation.
A single enzymatic system efficiently converts CO2 into CO and HCOOH through 2e− reactions. However, a single enzyme can efficiently and specifically convert or reduce CO2 to methane via an 8e− reaction. Yang et al.98 reported a CO2 reductase enzyme called remodeled nitrogenase. They substituted the Mo–Fe protein twice in the parent nitrogenase and described it as a new enzyme. When they used it for CO2 reduction, they observed that it has eight electron reduction capacities. Surprisingly, 1 nmol of the remodelled nitrogenase mediated the reduction of CO2 and produced 21 nmol of methane.
Different catalysts such as photocatalysts,99,100 electrocatalysts,101,102 and biocatalysts103–117 have been made to reduce CO2 to CH3OH, although the enzymatic system has been explored less for CO2 conversion to CH3OH through consecutive reduction catalyzed by three different dehydrogenases such as FDH, FaldDH and ADH in the presence of NADH. The first paper about this approach was published by Kuwabata et al.103 They used FDH and methanol dehydrogenase (MDH) with MV2+ as an electron mediator in an electrocatalytic cell. They found that HCOH was produced when the concentration of enzyme was kept low, whereas a mixture of HCOH and CH3OH was found when they increased the enzyme concentration. Later in 1999, Obert and Dave104 took a modified version of the enzymatic system and found methanol selectively and in a more efficient way. They have reported an efficient enzymatic system containing three dehydrogenases such as FDH, FaldDH and ADH instead of MDH as found by Kuwabata et al. in a silica sol gel matrix. When CO2 gas was bubbled into the solution containing the encapsulated enzyme together with NADH, they found CH3OH as an end product, as CO2 undergoes sequential reduction by the three different enzymes. Enzymes are generally unstable in an oxygen environment. Vigorous stirring of the enzymatic cell may have also contributed to the denaturation of the enzyme. So immobilization of the labile enzyme in support matrices, increases their stability, thus resulting in an enhanced probability of reaction because of an increase in the local concentration of the substrates within the nanopores and to some extent an increased product yield. Therefore, Obert and Dave immobilized the enzymes in a silica sol gel and found an increase in the CH3OH production compared to the free mobilized system.
In order to enhance the efficiency of the system, Wu et al.105 further modified the immobilization of ADH in the Obert and Dave enzymatic system. They adjusted the pore size of the silica gel in immobilization carrier (SC-PEG) by using poly(ethylene glycol) (PEG). Results showed that the PEG modification of the silica matrix increases the enzymatic activities to a smaller extent. Similarly, Xu et al.106 suggested another hydride system containing ALG–SiO2 gel encapsulated with dehydrogenase. This hybrid gel was prepared within the ALG solution where in situ growth of silica precursor was observed, followed by calcium ion (Ca2+) crosslinking. By using this hybrid gel they immobilized the three dehydrogenases, and found that the CH3OH production yield was highest for this ALG–SiO2 hybrid system (98.1%) than the pure ALG gel system. Furthermore, the CH3OH yield that was catalyzed by dehydrogenases in an ALG–SiO2 hybrid gel could be retained as high as 78.5% after recycling the gel 10 times and as high as 76.2% after 60 days storage. The significantly improved catalytic properties of the dehydrogenases in the ALG–SiO2 composite were attributed to the creation of the appropriate immobilizing microenvironment: high hydrophilicity, moderate rigidity, ideal diffusion characteristics, and optimized cage confinement effect.
A recent study by Sun et al. showed a green and efficient way of encapsulation.107 Through a simple and mild biomimetic mineralization process they encapsulated the three dehydrogenases, within titania (TiO2) particles and found that the amount of CH3OH extracted by each mole of enzymes was higher in comparison to the sol gel silica system. The activity of enzymes are preserved more effectively because of the biomimetic process that occurred at milder conditions and also the system is more favourable for CH3OH production because of the favourable interactions among enzymes are possible as TiO2 exhibits a nanoscale environment.107 In this method, all enzymes are encapsulated by co-immobilization, from which may arise the unnecessary collision between enzymes that cause the loss of its full potential. Luo et al.108 reported a sequential immobilization of different dehydrogenases. They sequentially immobilized the enzymes in a polymeric membrane by the fouling formation technique. The membrane consisted of a regenerated cellulose skin layer and a poly(propylene) (PP) non-woven support with an effective area of 13.4 cm2. The membrane support layer was positioned to face the feed in order to increase the enzyme loading and stability, and the skin layer was covered by an extra PP support to alleviate membrane compression and prevent peeling off. When CO2 was bubbled through, a high production of CH3OH was found compared to the free enzymatic and co-immobilization enzymatic system. In another approach, a novel capsule-in-bead scaffold for a spatially separated multi-enzyme system with layer-by-layer assembly and a biomineralization technique was proposed by Jiang et al.109 In this system they constructed a microcapsule of a guest molecule containing three dehydrogenases individually as shown in Fig. 11a. These guest dehydrogenase microcapsules were formed stepwise by using protamine, charged silica and calcium carbonate (CaCO3) sequentially. Then the capsule in-bead scaffold was made using co-encapsulation technology. This provides the temporal and spatial separation of enzymes which increased the CH3OH production yield as compared to the traditional co-immobilization as shown in Fig. 11b.
![]() | ||
Fig. 11 (a) Schematic representation of the preparation of capsule-in-bead for enzyme immobilization (b) plot of methanol production versus reaction time for the immobilized enzyme, free enzyme and co-immobilized enzyme (reprinted with permission from ref. 109). |
In most of the cases, the cofactor NADH was used and was not regenerated in the system. So, after a certain amount of time the process collapsed completely or in other words the process was continued until it ran out of NADH. However, natural cofactors are generally expensive and price has greatly hampered large scale operations. So a different system such as pure enzymatic,110,111 enzyme integrated photochemical,115,116 and electrochemical117 system was proposed for effective in situ regeneration of cofactor NADH.
For the CO2 reduction system, first a regenerating NADH system was proposed by El-Zahab et al.110 using an extra enzyme, GDH for cofactor regeneration. Here, three CO2 reducing enzymes were co-immobilized by using poly(styrene) particles together with GDHs as shown in Fig. 12. A suspension solution of the particle attached enzymes and cofactor were prepared, and then CO2 was bubbled through it. It was observed that the effective collision among the particles afforded sufficient interactions between the cofactor and enzymes, and thus enabled the sequential transformation of CO2 to methanol together with cofactor regeneration. It was seen that 0.02 μmol h−1 g−1 enzyme of methanol was obtained and 80% of the enzymes had retained their productivity after 11 reuse cycles.
![]() | ||
Fig. 12 Chemical route of enzymatic synthesis of methanol from CO2 with in situ regeneration of NADH (reprinted with permission from ref. 110, license no. 3845750833698). |
In continuing this work Ji et al.111 also used GDH, as an enzyme for cofactor generation. The cofactor and four enzymes including FDH, FaldDH, ADH, and GDH were co-encapsulated in situ, inside the lumen of hollow nanofibers by including them in the core-phase solution for coaxial electrospinning, in which the cationic polyelectrolyte was dissolved. The polyelectrolyte penetrating across the shell of the hollow nanofibers enabled efficient tethering and retention of cofactor inside the lumen via ion exchange interactions between the oppositely charged polyelectrolytes and cofactor as shown in Fig. 13. Here another enzyme, carbonic anhydrase, assembled on the outer surface of the hollow nanofibers and was used for accelerating the hydration of CO2. This five-enzyme-cofactor catalyst system exhibited high activity for methanol synthesis. The use of the hollow nanofiber support system produced a high amount, up to 103.2%, of methanol (the highest reported value so far), compared to the use of the free enzyme, which only produced 36.17%.
![]() | ||
Fig. 13 Diagrammatic representation of the set-up for co-axial electrospinning for the construction of a hollow nanofibre supported multienzyme for CO2 reduction. Here GDH is used as an NADH regenerating system (reprinted with permission from ref. 111. Copyright American Chemical Society 2015). |
Phosphite dehydrogenase (PTDH) can also be used as regenerating system for NADH. Cazelles et al.112 published a paper in which they describe using three different regenerating systems (see Fig. 14) such as PTDH, glycerol dehydrogenase (GlyDH) and a natural photosystem extracted from spinach leaves (chloroplast). PTDH was proven to be more efficient than others. They carried on the experiments by integrating a NADH regenerating enzyme with three dehydrogenases encapsulated in silica nanocapsules using phospholipids. This hybrid system shows an activity 55 times higher than that of the free enzymes.
![]() | ||
Fig. 14 Schematic illustration of multi-enzymatic system together with three enzymatic regeneration systems for NADH (reprinted with permission from ref. 112). |
In the case of photochemical enzymatic regeneration system, the photocatalyst is integrated for effective regeneration of NADH. Liu et al.113 reported a carbon nitride array (s) material, which when combined with a three dehydrogenases system gives an efficient methanol production yield together with NADH regeneration. Here the CNA was prepared using a sacrificial diatom template and its scanning electron microscopy (SEM) images are shown in Fig. 15a. By using CNA integrated with enzyme and NADH (which act as a template for enzyme immobilization as well as the NADH regenerating system), Rh(III) complex (mediator), and TEOA as a sacrificial electron donor, methanol was found as a product (0.21 mmol methanol min−1 per g enzyme) under visible light irradiation as shown in Fig. 15b.
![]() | ||
Fig. 15 (a) SEM images of CN and (b) the electron flow in the photocatalytic system of NAD+ reduction composed of a robust CNA photocatalyst, Rh(III) electron mediator and TEOA as a sacrificial reagent (reprinted with permission from ref. 113). |
Later Aresta et al.114 extensively studied the photochemical regeneration of NADH by using different photocatalysts such as copper(I) oxide (Cu2O), indium vanadate (InVO4), rutin@TiO2, and chromium-modified TiO2 {[CrF5(H2O)]2−@TiO2}. They found that [CrF5(H2O)]2−@TiO2 gave a good yield of NADH regeneration in the presence of a Rh bipyridine complex in a watery bioglycerol mixture upon irradiation of visible light. This catalyst was seen to be an efficient and selective reduction catalyst for NAD+ to 1,4-NADH. Upon irradiation of light, the photocatalysts generate an exciton and these excited electrons were moved by the mediator to the surface of the photocatalyst, where they carried out the reduction of NAD+ to NADH, involving the electron mediator and the oxidation of glycerol, and simultaneously CO2 was reduced to methanol. But for [CrF5(H2O)]2−@TiO2, the mechanism showed that the photoexcited electrons were transferred from the Cr-complex to the conduction band of TiO2, then these electrons were moved to the Rh complex which acted as a electron transferring agent. It was found that the reduction of the pentamethylcyclopentadienyl rhodium 2,2′-bipyridin {[Cp*Rh(bpy)·(H2O)]2+} complex to [Cp*Rh(bpy)] adds a proton and results in the conversion into a hydrido form, which further donates hydride and electrons to NAD+ for reduction. Because of this reason, the [CrF5(H2O)]2−@TiO2 exhibits higher activity than other photocatalysts. Dibenedetto et al.115 suggested another photocatalyst, ZnS-A, which can effectively regenerate NADH in an enzymatic system upon irradiation of light.
Similarly, Yadav et al.116 reported a graphene-based photocatalyst with sequentially coupled enzymes. They first used CCGCMAQSP, but they found a very low conversion rate. For improved performance they took a combination of two highly versatile chromophoric motifs, isatin and porphyrin together with chemically converted graphene which is collectively known as robust CCG-IP. Isatin and 2-aminoanthraquinone exhibited a much lower absorption band between 400 and 500 nm. So they concluded that CCG-IP is capable of efficient visible light harvesting and thus provides photocatalytic energy for the highly efficient regeneration of enzymatically active NADH, which is vital for methanol formation. According to Yadav et al., in the electron transfer mechanism for the whole system, as shown in Fig. 16a, the photoexcitation electrons are generated at IP and these electrons transfer to CCG and then to the Rh complex. Addo et al.117 proposed another system in which poly(neutral red) is deposited on a glassy carbon electrode (GCE) which efficiently regenerates NADH from NAD+ in a multi-enzymatic system as shown in the Fig. 16b. In this system, carbonic anhydrase was also used for more hydration of CO2 for efficient reduction to methanol.
![]() | ||
Fig. 16 (a) The schematic electron flow in the photocatalytic system of NAD+ reduction composed of a robust CCG-IP photocatalyst, Rh(III)-electron mediator and TEOA as a sacrificial reagent and the methanol production by a multi-enzymatic system (reprinted with permission from ref. 116. Copyright American Chemical Society 2014). (b) Schematic representation of multi-enzymatic CO2 reduction to methanol and here NADH regeneration is done with a GCE (reproduced from ref. 117). |
![]() | ||
Fig. 17 Assembly of an artificial photosystem such as light harvesting antenna, photosensitizer, electron donor/ acceptor, oxidation catalyst and reduction catalyst (reprinted with permission from ref. 119). |
Enzyme | Process | Materials | Immobilization support | Product | Catalytic efficiency/product yield | Ref. |
---|---|---|---|---|---|---|
a BABM = bioadhesion-assisted bio-inspired mineralization; CCGIP = chemically converted graphene–isatin porphyrin; MgATP: magnesium adenosine triphosphate; p-InP = p-type-indium phosphide; PAH = polycyclic aromatic hydrocarbons; PU = poly(urethane); SrTiO3 = strontium titanate. | ||||||
CODH M. thermoacetica | Electrocatalytic | GCE, MV2+ | — | CO | 700 h−1 | 52 |
(Ch)CODH | Electrocatalytic | Pyrolytic graphitic edge | — | CO | — | 54 |
(Ch)CODHI | Photoelectrocatalytic | PGE electrode, N-type semiconductor (CdS/TiO2), MES | — | CO | — | 55 |
(Ch)CODHI | Photocatalytic integrated enzyme | TiO2, RuP sensitizer | — | CO | 530 h−1 | 56 |
(Ch)CODHI | Photocatalytic integrated enzyme | TiO2 (P-25)/ZnO/SrTiO3, RuP sensitizer, MES | — | CO | 0.14 s−1 | 51 |
(Ch)CODHI | Photocatalytic integrated enzyme | CdS QDs, MES | CO | 1.23 s−1 | 57 | |
FDH P. oxalaticus | Enzymatic | NADH | — | HCOOH | 3 s−1 | 62 |
FDH | Photoelectrocatalytic | p-InP, MV2+ | — | HCOOH | 80–93% | 63 |
TbsFDH | Enzymatic | Sodium phosphate buffer, NADH | — | HCOOH | 5.8-fold more than CbFDH | 64 |
FDH | Enzymatic | NADH, Tris–HCl buffer | ALG–SiO2 | HCOOH | 95.6% | 65 |
CbFDH | Enzymatic | NADH, TEOA, phosphate buffer, carbonic anhydrase | — | HCOOH | 2.17 × 103 μmol min−1 | 66 |
FDH | Photocatalytic integrated enzyme | ZnTMPyP, MV2+, TEOA | — | HCOOH | 62 μmol/3 h | 67 |
FDH | Photocatalytic integrated enzyme | NADPH, MV2+, chlorin-e6, bis–Tris buffer | Silica gel | HCOOH | 20 μmol/3 h | 68 |
CbFDH | Photocatalytic integrated enzyme | ZnTPPS, NADPH, NADH, Pt NPs, MV2+, GDH | — | H2, HCOOH | 0.2 μmol/1.5 h | 69 |
FDH | Photocatalytic integrated enzyme | NADH, Rh sensitizer, graphene-based photocatalyst | — | HCOOH | — | 70 and 71 |
Mo-FDH R. capsulatus | Enzymatic | NADH, potassium phosphate buffer | — | HCOOH | 1.5 s−1 | 73 |
W-FDH (1 and 2), S. fumaroxidans | Enzymatic | MV+, sodium bicarbonate | — | HCOOH | 2.5 × 103 s−1 | 76 |
0.2 × 103 s−1 | ||||||
W-FDH1 S. fumaroxidans | Electrocatalytic | PGE electrode, buffer | — | HCOOH | 0.5 × 103 s−1 | 77 |
W-FDH C. carboxidivorans | Enzymatic | NADH, NADPH, phosphate buffer | — | HCOOH | Kcat = 0.08 | 79 |
FDH A. Woodii | Enzymatic | H2, buffer | — | HCOOH | 28 s−1 | 74 |
EcFDH-H E. coli | Electrocatalytic | MV2+, MES, graphite-epoxy electrode | — | HCOOH | <1 s−1 | 75 |
Remolded nitrogenase | Enzymatic | Buffer containing sodium dithionate and MgATP regenerating system | — | CH4 | 21 nmol | 98 |
FDH and FaldDH | Enzymatic-BABM approach | NADH | TiO2 NPs | HCOH | 80.9% | 82 |
FDH and FaldDH | Enzymatic | NADH, phosphate buffer | — | HCOH | — | 83 |
ADH | Photocatalytic integrated enzyme | ZnTPPS, MV2+, NAD+, TEOA, NADH | — | CH3OH | Conversion ratio 2.3% | 84 |
FDH, MDH | Electrocatalytic | MV2+ | — | CH3OH | 0.05 μmol h−1 | 103 |
FDH, FaldDH, ADH | Enzymatic | NADH | Silica sol gel matrix | CH3OH | 91.2% | 104 |
FDH, FaldDH, ADH | Enzymatic | NADH | SC-PEG | CH3OH | — | 105 |
FDH, FaldDH, ADH | Enzymatic | NADH | ALG–SiO2 | CH3OH | 78.5% | 106 |
FDH, FaldDH, ADH | Enzymatic-biomimetic mineralization process | NADH | TiO2 | CH3OH | 2.1 μmol h−1 | 107 |
FDH, FaldDH, ADH | Enzymatic | NADH, Tris–HCl buffer | Cellulose–PP | CH3OH | 0.7 mM/1.5 h | 108 |
FDH, FaldDH, ADH | Enzymatic-layer-by-layer biomineralization | NADH | Protamine-charged silica–CaCO3 | CH3OH | 1.5 μmol h−1 | 109 |
FDH, FaldDH, ADH, | GDH, NADH | Polystyrene | CH3OH | 0.02 μmol h−1 g−1 | 110 | |
FDH, FaldDH, ADH | Enzymatic | NADH, GDH, carbonic anhydrase | Hollow PU nanofibres, PAH, glutamic acid | CH3OH | 36.17% | 111 |
FDH, FaldDH, ADH | Enzymatic | PTDH, GlyDH, chloroplast | Silica nanocapsules by phospholipids | CH3OH | — | 112 |
FDH, FaldDH, ADH | Photocatalytic integrated enzyme | Rh complex, TEOA | CNA | CH3OH | 0.21 mmol h−1 g−1 | 113 |
FDH, FaldDH, ADH | Photocatalytic integrated enzyme | NADH, Cu2O, InVO4, rutin@TiO2, [CrF5(H2O)]2−@TiO2, NADH, Rh complex, bioglycerol–water | Ca alginate–TEOS | CH3OH | — | 114 |
FDH, FaldDH, ADH | Photocatalytic integrated enzyme | ZnS-A, NADH, sodium dithionite, Tris–HCl buffer, bioglycerol | Ca alginate–TEOS | CH3OH | — | 115 |
FDH, FaldDH, ADH | Photocatalytic integrated enzyme | NADH, CCGIP, Rh complex, TEOA | — | CH3OH | 11.21 μM | 116 |
FDH, FaldDH, ADH | Electrocatalytic | GCE, neutral red, NADH, carbonic anhydrase, phosphate buffer | — | CH3OH | 0.6 mmol g−1 h−1 | 117 |
(a) The enzymes used for this are very sensitive and are unsuitable for long-term use. They easily get denaturised, harsh conditions, even intense sunlight, can decrease its activity, but the major drawback is the costly protein purification process through which the enzymes are extracted. These negatives hinder this enzymatic system for large scale use.
(b) For an industrial exploitation, the design of a suitable bioreactor is a key issue. A solution needs to be found for several important issues such as the discovery of efficient systems for solar energy harvesting, a system for easy electron transfer to CO2 for an efficient reduction, proficient charge separation systems, space separation of oxidation and reduction processes on the catalysts, selective synthesis of hydrocarbon, and use of low cost compounds for catalyst design.
Fortunately, several approaches and systematic studies are in progress to make “enzymatic conversion of CO2” a sustainable and effective one. A de novo catalyst has been proposed by mimicking the nature of the enzymes so that costly enzymes can be replaced. Enzymes whose active site is composed of metal ions (such as Mn, W, Mo, Fe, and Ni) in a specific rearrangement of amino acids provides a way to mimic it in an efficient manner. Many researchers have reported systems where PSII, hydrogenase enzymes are mimicked in light for water splitting. However, very little work has done relating to mimicking CO2 reducing enzymes. Furthermore, computational study using ab initio and molecular simulation can provide a deeper understanding of active sites and reaction mechanisms. Finally, the solution of such problems will require time, but all targets can be reached as there is no real (negative thermodynamics, for example) barrier. The list of abbreviations used in the review is presented in Table 2.
Abbreviations | Full form |
---|---|
ADH | Alcohol dehydrogenase |
ALG | Alginate |
CbFDH | Candida boidinii formate dehydrogenase |
CCG-BODIPY | Chemically converted graphene 1-picolylamine-2-aminophenyl-3-oxy-phenyl-4,4′-difluoro-1,3,5,7-tetramethyl-2,6-diethyl-4-bora-3a, 4a-diaza-S-indacene-triazene |
CCGIP | Chemically converted graphene–isatin porphyrin |
CCGMAQSP | Chemically converted graphene multi-anthraquinone substituted porphyrin |
CNA | Carbon nitride array |
CODH | Carbon monoxide dehydrogenase |
DME | Dimethyl ether |
F-Gases | Fluorinated gases |
FaldDH | Formaldehyde dehydrogenase |
FDH | Formate dehydrogenase |
GDH | Glutamate dehydrogenase |
GHG | Greenhouse gas |
GlyDH | Glycerol dehydrogenase |
IPCC | Intergovernmental panel on climate change |
LUMO | Lowest unoccupied molecular orbital |
MDH | Methanol dehydrogenase |
MES | 2-(N-Morpholino)ethanesulfonic acid |
MV | Methyl viologene |
NADH | Nicotinamide adenine dinucleotide-H2 |
NADPH | Nicotinamide adenine dinucleotide phosphate-H2 |
NPs | Nanoparticles |
PCET | Proton coupled electron transfer reaction |
PEG | Poly(ethylene glycol) |
PGE | Pyrolytic graphite edge electrode |
PTDH | Phosphite dehydrogenase |
TbsFDH | Thiobacillus formate dehydrogenase |
TEOA | Triethanolamine |
YADH | Yeast alcohol dehydrogenase |
ZnTMPyP | Zinc tetrakis(4-methylpyridyl)porphyrin |
ZnTPPS | Zinc tetraphenylporphyrin tetrasulfonate |
This journal is © The Royal Society of Chemistry 2016 |