Journal of Materials Chemistry A View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. REVIEW View Journal Carbon dioxide conversion to synthetic fuels using biocatalytic electrodes Cite this: DOI: 10.1039/c6ta07571a Stefanie Schlager,*a Anita Fuchsbauer,b Marianne Haberbauer,c Helmut Neugebauera and Niyazi Serdar Sariciftci*a Carbon dioxide has evolved from being considered as a greenhouse gas to a valuable carbon feedstock for the generation of artificial fuels and valuable chemicals. In this work, we review the biocatalytic approaches towards CO2 conversion into chemicals and fuels. We display the opportunities and challenges of using biocatalysts. Our work especially focuses on bio-electrocatalytic systems. These electrochemical applications of biocatalysts gain increasing interest, as electrochemical redox processes can avoid Received 1st September 2016 Accepted 2nd December 2016 expensive mediators and co-factors. This is also a pathway for renewable energy storage because wind or solar energy can possibly be applied as electrical sources for the electrochemical CO2 conversion DOI: 10.1039/c6ta07571a www.rsc.org/MaterialsA systems. Biocatalytic CO2 conversion together with renewable energy storage represents a viable and sustainable route for the generation of chemicals and fuels. 1. Introduction In recent years carbon dioxide has been regarded as the major greenhouse gas, partly responsible for the global climate change since greenhouse gas emissions are increasing and fossil fuels are depleting increasingly.1,2 As a result of burning hydrocarbons, carbon dioxide is released to the atmosphere to an ever increasing extent. Although it is reported that methane and water vapor contribute even more to atmospheric greenhouse gases, CO2 has a much higher impact on global warming due to a much longer residence time in the atmosphere.3,4 This duration is still just estimated for CO2 but it is at least 3 times higher in comparison to that of methane and 7 times higher than that of water vapor. In terms of its concentration in the atmosphere, CO2 is the second most abundant greenhouse gas next to water vapor. This linear transformation of the fossil fuel bound carbon into atmospheric carbon dioxide is to be changed to a cyclic use of carbon dioxide. This would bring the necessity of capturing and recycling of CO2. With this strategy of cyclic use, carbon dioxide can be regarded as carbon feedstock for future generation of fuels and chemicals. CO2 neutral and renewable fuel generation and use can bring our atmosphere to a sustainable and stabilized equilibrium. In contrast to the techniques of Carbon Capture and Sequestration (CCS), where CO2 is stored in cavities under sea and land, Carbon Capture and Utilization (CCU) uses carbon a Linz Institute for Organic Solar Cells (LIOS), Johannes Kepler University Linz, Altenbergerstraße 69, 4040 Linz, Austria. E-mail: [email protected] b c PROFACTOR, GmbH, Steyr, Austria dioxide as a valuable carbon source for a variety of chemicals, materials and fuels.5–8 Nevertheless, for products generated from CO2 as a carbon source, it is required to chemically reduce carbon dioxide e.g. to carbon monoxide or to hydrocarbons. Since the carbon dioxide molecule is a very stable and low energy conguration of carbon, such reduction reactions need a lot of energy input (1st and 2nd law of thermodynamics) as well as cleverly designed catalysts to overcome the energy barriers in such reactions. Besides numerous studies on chemical and electrochemical techniques using synthetic routes and synthetic catalysts, we here focus on the biochemical and bio-electrochemical routes toward CO2 conversion and biofuel generation.9 In these biocatalytic approaches we use natural processes as a model for the choice of catalysts and process design. Biocatalytic as well as bio-electrocatalytic approaches have several advantages in comparison to routes using synthetic catalysts. From the comparison of synthetic catalysts with biocatalysts it is obvious that the source or the synthetic pathway of the catalyst plays an important role. While catalysts, such as organic molecules or metal–organic complexes, have to be synthesized prior to their use and therefore oen require elaborate steps and materials, biocatalysts can be obtained directly from natural sources and can be provided in high amounts. However, probably the most noteworthy properties of biocatalysts, like microorganisms or enzymes, are high selectivity and yield with regard to the desired product. Moreover such processes can be tuned to various possible products by choosing the biocatalysts accordingly. Another advantageous property to mention is that reactions can be performed under rather mild reaction conditions (atmospheric pressure and ambient temperature). Bio-electrocatalytic applications furthermore offer Acib, GmbH, Linz, Austria This journal is © The Royal Society of Chemistry 2017 J. Mater. Chem. A View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Journal of Materials Chemistry A the possibility of direct electron injection into the biocatalysts where no mediators or electron equivalents are required. In addition such techniques represent a highly attractive pathway for chemical storage for renewable energies, because they can serve as an electrical energy source. In the following section different biocatalytic approaches for CO2 conversion are discussed. In general we distinguish between living biocatalytic systems such as microorganisms and non-living biocatalysts such as enzymes. Furthermore we comparatively present studies on biocatalytic as well as bioelectrocatalytic processes.10 2. Microorganisms as biocatalysts for CO2 conversion 2.1 Conversion using reduction equivalents Microorganisms particularly gained interest in carbon capture and utilization research due to the ability to convert CO2 to a broad range of possible valuable products and fuels. Application of such microorganisms has become highly attractive as several different strains of pure as well as mixed cultures of microorganisms are suitable for application in biofuel and biochemical generation.11 Utilization of anaerobic bacteria has since become attractive, as they provide suitable conditions for biotechnological processes and biofuel production. Several studies showed that such microorganisms can be favorable for chemical and biofuel generation from fermentation. Particularly the generation of acetate or ethanol has been a main focus.12 Those products are obtained due to the metabolism of anaerobic bacteria as hydrocarbons serve as nutrients. However, also the viability to grow acetogens using H2 and CO2 or CO instead of glucose was shown. Acetate formation was also displayed for different growing conditions. Moreover various models for the role of enzymes and possible metabolism pathways for acetogenesis have been proposed.13 Besides high selectivity towards a certain product, microorganisms are also favorable in terms of representing a selfregenerating catalyst, long-term stability, sustainability and biocompatibility. In addition, sources of the microorganisms are diverse and highly abundant.14 By tuning the kind of microorganisms and/or environmental conditions such as temperature and pressure, the metabolism of the biocatalysts can be directed toward one or even several desired products.15 This direct conversion of CO2 using microorganisms opens a favorable way for the generation of biofuels and biochemicals in terms of ethical issues as well. In contrast to biofuels from crop or other biomass which may compete with food production, CO2 can be converted directly using microorganisms and serves as the only carbon source. This microorganism-catalyzedmethod is therefore not competing with food supply since biofuels and biochemicals can be obtained from a non-food source. However, for achieving high yield and high selectivity using this approach, it is required to understand the corresponding enzymology and mechanism of the metabolism.16,17 To compete with fossil fuels, biofuels need to have high energy density. Therefore, the main interest is on liquid target J. Mater. Chem. A Review materials for the direct generation of e.g. higher alcohols such as ethanol or butanol. Liou et al. demonstrated the generation of e.g. acetate, ethanol and even butanol from CO, CO2 and sugars as a growth support and carbon sources for an anaerobic Clostridium strain. Those strains were chosen because the Wood–Ljungdahl pathway for Clostridia serves as a model route. This path gained main interest as it consists of multiple redox steps, each catalyzed by a corresponding enzyme, and therefore provides generation of various possible products.18,19 Detailed investigation of the Wood–Ljungdahl pathway as a result led the focus toward application of archaea or bacteria such as Clostridia.20 Younesi et al. focused their research on the generation of acetate and ethanol as fuels from Clostridium ljungdahlii. In their study they investigated the bioconversion of syngas containing CO, CO2 and H2. Besides the inuence of pressure they also studied growth inhibition and ethanol yield upon CO content. They found acetate as the major product and in general enhanced production rates of both, ethanol and acetate, when the pressure was increased to 1.8 atm.21 Also the groups of Ragsdale and Tracy furthermore presented studies based on investigations of the Wood–Ljungdahl pathway. In contrast to the group of Younesi, who investigated syngas conversion by acetogens, Ragsdale et al. examined CO2 xation or conversion and the corresponding enzymology.22,23 A detailed study on the conversion of CO2 and other substrates for biofuel generation and biorenery application was carried out by the group of Tracy. From investigations of the enzymatic mechanism in microorganisms' metabolism according to the Wood–Ljungdahl pathway, dependence of the desired products on the source of substrate (e.g. CO2, cellulose, glycerol) was observed (Fig. 1).24 Schiel-Bengelsdorf and Dürre moreover presented a review of several acetogens following the Wood–Ljungdahl pathway for the generation of acetate and advanced products like ethanol and butanol (Fig. 2).25 Results on the potential of these microorganisms toward CO2 xation and conversion gave rise to intensive research toward syngas conversion with CO2 as the carbon source for biochemicals and biofuels. Munasinghe et al. moreover reviewed studies on syngas fermentation to biofuels using Clostridia and acetogenic microorganisms.26 In another study the potential of microbial processes toward syngas conversion, in comparison to the conversion of sugar, toward different products using Moorella thermoacetica was displayed by the group of Hu et al. They found productivity rates at 0.585 g L 1 h for acetate only, which is in contrast to acetogenic Clostridium ljungdahlii, as presented e.g. by the groups of Younesi, who also found ethanol as a second product.21,27 Particularly reaction conditions play an important role in microbial processes for the synthesis of biofuels like ethanol or higher alcohols like propanol and butanol. Besides parameters like temperature, pH and pressure the constitution of buffer or medium solution has an enormous impact on not only the performance of syngas but also process costs.28,29 All these approaches showed the broad variety of possible biofuel and biochemical generation from syngas or other substrates to be converted. Above all, however, especially biosynthesis using gaseous CO2 directly as a carbon source This journal is © The Royal Society of Chemistry 2017 View Article Online Journal of Materials Chemistry A Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Review Wood–Ljungdahl pathway for the conversion of different carbon based substrates showing the individual steps with the corresponding enzymes. Figure reproduced with permission from ref. 24. Fig. 1 gained high interest in this eld. In general the strategy is to nd a combination of generating biofuels and biochemicals, reducing net carbon dioxide emission and nding a carbonneutral, renewable system solution.30–32 For such a direct conversion of CO2 Deppenmeier et al. published a detailed study on using methanogenic microorganisms. In their paper they show the stepwise mechanisms of the metabolism with the corresponding enzymes for the conversion of CO2 together with H2 to methane, CH4. They especially focused on microorganism strains belonging to This journal is © The Royal Society of Chemistry 2017 Methanosarcina for investigating redox processes with electron and hydrogen transfers involved. They propose that the process of methane formation is due to proton and ion gradients in the microorganisms that are generated from participating enzymes and proteins over the membrane.33 Carbon xation pathways in general can be divided into six different possibilities. The rst approach is the pathway via the Calvin cycle or reductive pentose phosphate pathway from photosynthesis and is normally used by algae, plants and cyanobacteria. In this cycle CO2 is xed through carboxylation of J. Mater. Chem. A View Article Online Review Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Journal of Materials Chemistry A Possible acetogenic microorganisms with their optimum growth conditions and obtained products. Figure reproduced from SchielBengelsdorf et al.25 Fig. 2 D-ribulose-1,5-diphosphate and subsequent conversion into two molecules of 3-D-phosphoglycerate. Alternatively there are ve possible autotrophic pathways where carbon xation is performed due to incorporation of CO2 into a carbon backbone and using acetyl-CoA and/or succinylCoA cycles for carbon xation. A detailed description of the possible CO2 xation pathways has been presented by Ducat and Silver.34 However, these strategies provide information J. Mater. Chem. A about molecules obtained via carboxylation reactions. A direct conversion of CO2 to acetate, ethanol or other fuels is only obtained from the Wood–Ljungdahl route as displayed in Fig. 1. The Wood–Ljungdahl pathway for microorganisms, in particular for Clostridia, exhibits their favorable properties, as it is possible to obtain several products. Besides higher alcohols like butanol, requiring a multi-step cascade of corresponding enzymes, also shorter pathways are possible. From products This journal is © The Royal Society of Chemistry 2017 View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Review like acetate moreover methane generation can be obtained. Depending on the strain, products can be chosen from a huge variety. In addition, considering reaction conditions, these products can be obtained at high yields and selectivities, making microorganisms notably attractive. Among the 83 species of methanogens described so far (including six synonymous species) 61 species (including ve synonymous species) are capable of oxidizing H2 and reducing CO2 to form methane. Examples of hydrogenotrophs are Methanosarcina barkeri, Methanosarcina thermophila, Methanococcus thermolithotrophicus, Methanocaldococcus jannaschii, Methanospirillum hungatei, Methanobacterium palustre, Methanobacterium formicicum or Methanothermus fervidus.35 To obtain high selectivity and yield there have been also approaches in genetic modication and synthetic biology. These studies aim at designing systems that meet the requirements of high quality biofuels that are further generated from sustainable sources and approaching a carbon neutral pathway at the same time.36–38 Further there have also been industrial applications and pilot processes that already implemented such technologies for carbon dioxide conversion using e.g. syngas fermentation processes or steel ue gas and microorganisms as catalysts.39–41 However, all of these approaches require hydrogen as a reduction equivalent and/or supporting mediators for charge and proton transport for the conversion of gaseous CO2.42 Hydrogen as well as mediators have to be produced previously e.g. in case of H2 from water splitting which causes a further costintensive step in addition to the actual CO2 conversion and biofuel generation. For this reason, focus is increasingly on substitution or even avoidance of such supplements. For this purpose the technique of microbial electrolysis evolved. Microbial electrolysis uses microorganisms attached to electrodes. This is possible because some microorganisms can be addressed directly in electrochemical systems for application in redox reactions and therefore additional mediators are not required anymore. Moreover, if renewable energies like solar or wind can be used for such bio-electrochemical conversion systems, those processes may also be considered as an opportunity for chemical energy storage. The time and place of generation of renewable energies are not aligned with the time and place of demand and consumption. A combination of a microbial system for reducing the greenhouse gas CO2 together with an electrochemical system, driven by a renewable energy source, would meet the requirements of both renewable energy storage and sustainable biosynthesis at once (Fig. 3).43,44 2.2 Journal of Materials Chemistry A electrosynthesis and by bio-electrocatalytic conversion of CO2. However, probably the most crucial part for such methods is charge transfer to and from biocatalysts and to understand the mechanism of the redox processes. In living biocatalysts like microorganisms charge transfer may occur outside the cell via the outer membrane. This is different from catalysts like the widely used metal–organic compounds, which also comprise non-living biocatalysts like enzymes, where charge transfer occurs via conjugated bonds and metal ions. Electrochemical reduction reactions in microorganisms are reported to be possible in the so-called exoelectrogenic strains, capable of transferring charges over the cell membrane. These transfers usually happen via extracellular electron transport by two possible mechanisms: direct electron transfer and mediated electron transfer.47–49 Rosenbaum et al. screened the different possibilities for such extracellular electron transfer mechanisms in the case of biocathodic microorganisms. For the direct electron transfer, electron-shuttling via cytochromes (proteins that enable redox reactions) on the outer membrane of the cell is proposed.50 The involved c-type cytochrome electron transfer chains can further be assisted by hydrogenases (cytochrome-hydrogenase partnership).51 Direct electron transfer furthermore can be performed with other enzymes, similar to a c-type cytochrome in different biochemical processes (Fig. 4).52,53 In the case of mediated electron transfer mediators can act endogenously (in the cell) or exogenously (outside the cell). Electron mediators such as neutral red and methylviologen are widely used for microorganisms as well as for non-living biocatalysts such as dehydrogenase enzymes.54–57 In a different investigation in terms of charge transfer in microorganisms, the group of Ajo-Franklin examined the Microbial electrosynthesis Application of microorganisms in electrochemical cells was found to be a favorable step toward substitution of mediators and additives. Production, supply and regeneration of mediators and other additives pose the problem of high costs and energy inputs. This is due to steps and processes that are necessary in addition to the actual CO2 conversion.45,46 Many of those extra steps can be avoided using microbial This journal is © The Royal Society of Chemistry 2017 Fig. 3 Correlation of time and space according to utilization and transport for renewable energies. J. Mater. Chem. A View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Journal of Materials Chemistry A electron transfer between living and non-living organisms. From an understanding of such electrical conduction in living organisms they could generate a synthetic electron conduit and therefore tune the property of a certain strain according to their redox capabilities.58 Further they tried to apply nanostructures for an even improved charge transport via cell membranes.59 All of these studies give advanced insight into the mechanistic pathway of redox reactions and charge transfer in biocatalytic and bio-electrocatalytic systems using microorganisms. This is Review crucial for the application for biosynthesis and improvement of bio-electrocatalytic processes. The rst report on bioelectrochemically synthesized methane from CO2 without any electron shuttle or mediators was published by Cheng et al. They showed methane generation from CO2 or carbonate conversion. Besides the generation of hydrogen from organic matter using exoelectrogenic bacteria, also an evolving methane concentration due to the presence of carbonate in the electrolyte solution was displayed. Methane generation actually was an unwanted side effect. However, they were one of the rst groups who moreover observed favored methanogenesis when hydrogen was apparent to a high extent.60,61 In a subsequent study Cheng and co-workers further found for the rst time that methane could be generated from electrical current by direct electron injection into microorganisms grown on a cathode instead of methane generation from hydrogen gas evolution, as presented previously. For their electromethanogenic approach they presented electron capture efficiencies of 96%.62 The potential to use such microbial electrolysis cells for biofuel generation was further addressed by Rabaey et al. They investigated methane generation from hydrogenophilic microorganisms and combined methanogenesis with anaerobic digestion. They used the oxidation of acetate to donate electrons for the reduction reaction. Together with cathodically generated hydrogen, wastewater was then converted to methane.63,64 Related to these previous studies Villano et al. proposed in their paper a combined process for the bioelectrochemical reduction of CO2 to methane in a bioelectrochemical system (BES) applying a methanogenic culture. Besides indirect reduction via abiotically generated H2 they also found direct reduction to methane due to extracellular electron transfer. At potentials more negative than 650 mV vs. SHE abiotic generation of hydrogen was observed. In addition to CO2 ushing, bicarbonate (NaHCO3) was present in solution as the buffer salt. Both sources of CO2 (as gas and as a bicarbonate salt) served as the carbon source to be reduced to methane by providing electrons directly from an electrode. In another study they further combined cathodic methane generation with a microbial anode to oxidize acetate to provide electrons.65,66 A similar approach to cathodically convert CO2 to methane was presented by Jiang et al. In contrast to Villano et al. they also Fig. 4 Proposed electron transfer mechanisms in microorganisms via (A) direct electron transfer involving c-cytochrome, (B) mediated electron transfer to a periplasmic hydrogenase or (C) direct electron transfer due to cytochrome-hydrogenase partnerships. Figure reproduced with permission from ref. 50. J. Mater. Chem. A Two compartment microbial electrolysis cell used by Zhen et al. for the electromethanogenesis with a biocathode. Figure reproduced with permission from ref. 70. Fig. 5 This journal is © The Royal Society of Chemistry 2017 View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Review observed acetate formation using potentials of 850 mV vs. Ag/AgCl (corresponds to 620 mV vs. SHE) and even more negative potentials. They found that acetate production increased when the potential was set to more negative values. Also, besides acetate, methane generation increased subsequently in comparison to more positive potentials. The coulombic efficiency of methane as the product increased to 95.8% at a set potential of 950 mV, whereas it decreased again to 56.7% at a more negative potential of 1150 mV. The coulombic efficiencies of acetate as the product uctuated in a range of 0–28.4% for set potentials between 950 mV and 1050 mV.67 Also Jourdin et al. observed production of acetic acid in anaerobic microbial electrosynthesis operated at 850 mV vs. SHE with bicarbonate as the carbon source. Differently, this group used microorganisms from planktonic cells. They found three predominant species of microorganisms in the biolm on their electrode. Acetoanaerobium from the order Clostridiales is known for the ability to convert CO2 to acetate and was mainly dominant in the biolm. Besides, Hydrogenophaga from the second apparent order, Burkholderiales, was dominant in suspension as well as in the biolm and was also expected to contribute to acetate formation. For microbial synthesis they conrmed from comparison experiments that 100% of the electrons consumed were used for acetate.68 Zhen et al. recently published a study on a two compartment microbial electrolysis cell focusing on the bioelectrosynthesis of methane. They showed results on the application of different potentials and the corresponding change in hydrogen and methane generation. For a constant potential of 900 mV vs. Ag/AgCl (corresponds to 670 mV vs. SHE) applied they observed the highest methane production and moderate hydrogen generation. For a potential as low as 600 mV vs. Ag/AgCl (corresponds to 370 mV vs. SHE), only methane generation was observed, although to a low extent. However, there was no hydrogen evolution observed, since a higher overpotential was expected to be required.69 Furthermore, they displayed the generation of methane in a MEC at a constant cathodic potential of 900 mV to Biocathode consisting of carbon felt. Microorganisms (white algae-like coverage) were grown on the carbon felt electrode for direct electron injection. Fig. 6 This journal is © The Royal Society of Chemistry 2017 Journal of Materials Chemistry A 1400 mV vs. Ag/AgCl ( 670 mV to 1170 mV vs. SHE) using graphite felt cathodes (Fig. 5).70 Another point in the application of microorganisms in electrochemical cells is the viability of regeneration and therefore the possibility of long-term performance. A long-term performance of 188 days of a microbial electrolysis cell was presented by the group of Van Eerten-Jansen, who investigated methane generation at potentiostatic conditions between 550 mV and 800 mV vs. NHE. As a carbon or CO2 source they added bicarbonate as well.71 Recent studies on microbial electrosynthesis have also been presented by Bajracharya et al. who used pure as well as mixed cultures for CO2 reduction. In their investigations they applied an assembly of graphite felt and stainless steel as the cathode.72 In our work, we used a direct approach for CO2 reduction with an even lower potential of 700 mV vs. Ag/AgCl ( 470 mV vs. SHE). Methanogenic microorganisms were grown on a carbon felt electrode to provide electrons directly (Fig. 6). In contrast to the previously mentioned studies, we did not use bicarbonate in the electrolyte solution. Instead of this we purged the cell with gaseous CO2 as the carbon source only. Further we long-term operated the microbial electrolysis cell for over one year (Fig. 7). We established a stable system that was even suitable for a continuous and constant conversion of CO2 to methane. During potentiostatic electrolysis we also found hydrogen generation as was observed earlier by the group of Zhen.70 From the comparison with blank measurements with a carbon felt electrode only, we found that hydrogen evolution was only there with the microorganisms. In total, a faradaic efficiency of 22% could be determined for the long-term performance cell.73 Tuning the operation conditions as well as the microorganism strains has shown that microbial electrolysis is highly suitable for generation of biofuels and biochemicals.74 A summary of the discussed approaches is presented in Table 1. High selectivities and yields are unique advantages of biocatalysts and highlight Fig. 7 Continuous production of H2 and CH4 during the long-term performance of a microbial electrolysis cell with gaseous CO2 as the carbon source only. The amounts of H2 and CH4 in the headspace of the cell produced during potentiostatic performance were detected using gas chromatography. The results show constant production volumes.73 J. Mater. Chem. A View Article Online Journal of Materials Chemistry A Table 1 Summary of different microbial electrosynthesis approaches with corresponding parameters Microbial electrosynthesis Mixed culture Mixed culture Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Review Substrate for conversion Product Electrode material Cathodic potential Reference Glucose, cellulose CO2 H2, methane, acetate Methane Carbon cloth with Pt catalyst Carbon cloth 0.3–0.8 V vs. SHE Cheng et al.60 Cheng et al.62 CO2, NaHCO3 H2, methane Carbon paper 0.7–1.2 V vs. Ag/AgCl (0.499–0.999 V vs. SHE) 0.65–0.9 V vs. SHE Villano et al.65 Hydrogenophilic methanogenic culture Hydrogenophilic methanogenic culture Mixed culture NaHCO3 H2, methane Graphite 0.85 V vs. SHE Villano et al.66 CO2, NaHCO3 Carbon felt CO2 Mixed culture CO2, NaHCO3 Methane Flexible multiwalled carbon nanotubes on reticulated vitreous carbon (NanoWeb-RVC) for Carbon stick 0.85–1.15 V vs. Ag/AgCl (0.653–0.953 V vs. SHE) 0.85 vs. SHE Jiang et al.67 Planktonic cells H2, methane, acetate Acetate Mixed culture CO2 Methane Carbon stick/graphite felt Mixed culture Clostridium ljungdahlii NaHCO3 NaHCO3 Methane Acetate, H2 Graphite felt Graphite felt/stainless steel Mixed culture NaHCO3 Graphite felt/stainless steel Sporomusa ovata CO2, NaHCO3 Clostridium ljungdahlii CO2 Moorella thermoacetica CO2 H2, methane, acetate Acetate, oxybutyrate Acetate, oxybutyrate, formate Acetate Sporomusa ovata CO2 Acetate Brewery wastewater sludge CO2, NaHCO3 Adapted mixed acetogens from brewery wastewater sludge Mixed culture domestic wastewater treatment plant sludge Enriched culture from bog sediment Mixed culture CO2, NaHCO3 H2, methane, acetate Acetate, H2 CO2 Acetate, H2 Carbon felt 0.9–1.1 V vs. Ag/AgCl (0.69–0.89 V vs. SHE) Bajracharya et al.72 CO2, NaHCO3 Acetate Carbon ber rod Bajracharya et al.72 CO2 Methane, H2 Carbon felt 0.6 V vs. Ag/AgCl (0.39 V vs. SHE) 0.7 vs. Ag/AgCl (0.49 V vs. SHE) Unpolished graphite stick Graphite stick Graphite stick Ni-nanowire coated graphite stick Graphite granules and graphite rods Graphite granules and graphite rods their attractiveness. As a result, not only mimicking of natural process but even direct application of biological materials such as biocatalysts gains a lot of interest in terms of renewable energies and CO2 reduction. Microorganisms offer high potential according to this purpose. However, considering the metabolic pathway of microorganisms, as discussed previously in this section, the reaction pathway of microorganisms consists of several, individual steps. Those steps are each catalyzed by a corresponding enzyme. The separate application of such single biocatalysts therefore additionally became attractive as certain reactions toward distinct products can be extracted from the multi-step pathway. This eases CO2 reduction reactions and makes it even more favorable as enzymes are not living J. Mater. Chem. A 0.6–1 V vs. Ag/AgCl (0.399 V vs. SHE) 0.9–1.4 V vs. Ag/AgCl (0.699–1.199 V vs. SHE) 0.7 V vs. NHE 0.6–0.9 V vs. Ag/AgCl (0.395–0.695 V vs. SHE) 0.6–1.1 V vs. Ag/AgCl (0.395–0.895 V vs. SHE) 0.6 V vs. Ag/AgCl (0.39 V vs. SHE) 0.6 V vs. Ag/AgCl (0.39 V vs. SHE) 0.6 V vs. Ag/AgCl (0.39 V vs. SHE) 0.6 V vs. Ag/AgCl (0.39 V vs. SHE) 0.79 V vs. Ag/AgCl (0.58 V vs. SHE) 0.79 V vs. Ag/AgCl (0.58 V vs. SHE) Jourdin et al.68 Zhen et al.69 Zhen et al.70 Van Eerten-Jansen et al.71 Bajracharya et al.72 Bajracharya et al.72 Bajracharya et al.72 Bajracharya et al.72 Bajracharya et al.72 Bajracharya et al.72 Bajracharya et al.72 Bajracharya et al.72 Schlager et al.73 biocatalysts but are able to operate at comparable yields and selectivities as living biocatalysts or microorganisms. In the following sections application of enzymes for the conversion of CO2 toward chemicals and biofuels will be discussed. 3. Enzymes as biocatalysts for CO2 conversion In the elds of renewable energies and CO2 conversion particular focus has been on the use of dehydrogenase enzymes as they are especially known for the ability to convert CO2 to carbon monoxide75,76 or hydrocarbons like formate, formaldehyde or methanol.77–81 Such dehydrogenases further occur in This journal is © The Royal Society of Chemistry 2017 View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Review Journal of Materials Chemistry A human liver where they perform in the opposite direction or oxidation to break down e.g. alcohols.82 Redox reactions catalyzed by dehydrogenases usually proceed under ambient conditions and high yields and are accompanied by the corresponding redox reactions of electron and proton shuttles such as the co-enzyme Nicotinamide adenine dinucleotide (NADH).83 Besides the application of dehydrogenase enzymes for the chemical reduction of CO2, other enzymes such as carbonic anhydrase are also capable of CO2 conversion. Carbonic anhydrase e.g. catalyzes the conversion of CO2 to bicarbonate in a hydrogenation reaction. Such reactions can also be combined with dehydrogenase processes to activate, enhance and accelerate CO2 reduction. Such improvements have been presented among others, e.g. by the group of Addo. They report on enhanced methanol generation from CO2 using dehydrogenase enzymes as the CO2 is converted to bicarbonate prior to the reduction reactions with carbonic anhydrase as the catalyst.57 Such improved mechanisms for electrofuel production from CO2 using carbonic anhydrase has also been described in detail by Hawkins et al.46 Other carboxylation reactions are e.g. the xation of CO2 as carboxyphosphate catalyzed by biotin carboxylase or the temporary xation by converting phosphoenolpyruvate to oxalacetate by phosphoenolpyruvate carboxylase. Similarly the enzyme ribulose-1,6-bisphosphatecarboxylase/oxygenase (RuBisCo) also catalyzes CO2 xation.84 This carboxylation step of ribulose-1,5-biphosphate is the initial step responsible for the ability of plants or algae to take up CO2.85 Another carboxylation, the one of 2-oxoglutarate to isocitrate, is catalyzed by isocitrate dehydrogenase.86 In general conversion to bicarbonate is ubiquitous and essential particularly in cells and reduction processes using living biocatalysts reported earlier in this review.87 However, the direct reduction of CO2 to fuels or valuable chemicals without any carbonation required prior to this, could be even more attractive as it spares an additional step. Such a direct conversion can be obtained by using the dehydrogenase enzymes carbon monoxide dehydrogenase, formate dehydrogenase, formaldehyde dehydrogenase and alcohol dehydrogenase. Carbon monoxide dehydrogenase (CODH) catalyzes the reduction of CO2 to CO. As a co-factor ferredoxin is used for providing electrons.80,88,89 However, using the enzymes formate-, formaldehyde- and alcohol dehydrogenase all together in a cascade is even more attractive, as methanol can be obtained and used directly as a fuel. The main advantage of utilizing isolated enzymes instead of directly using the microorganisms for certain reactions is the even higher selectivity. In comparison to living organisms, enzymes are restricted to mainly one product, independently from reaction conditions. A change in e.g. pH or temperature inuences mainly the performance and yield, but not the product itself.90 Furthermore handling and maintenance are more favorable in comparison to living biocatalysts as enzymes don't have to be kept alive by nourishing regularly. A well-known and common application of dehydrogenases, mainly formate dehydrogenase, is for catalyzing the oxidation of formate to CO2. Released electrons and protons further can be used for the regeneration of NADH to NAD+. This is crucial since many synthetic processes use NAD+ as electron and proton shuttle or reduction equivalent. Studies on regeneration of NADH or NAD+ respectively have been done e.g. by the group of Palmore and Whitesides. Whitesides et al. showed the application of formate dehydrogenase in the regeneration of NADH from its use in other synthetic routes. Palmore et al. later merged a dehydrogenase catalyzed redox process with benzylviologen catalyzed NAD+ regeneration in a methanol/ dioxygen fuel cell. In their process methanol was oxidized to CO2 with the aid of NAD+. The generated NADH was then regenerated to NAD+ by using benzylviologen as an electron provider.91,92 Also Neuhauser et al. investigated the regeneration of NADH by the conversion of formate to CO2 using formate dehydrogenase.93 However, besides the application of dehydrogenase enzymes for NADH regeneration, primarily interest evolved in the opposite reaction pathway, namely reduction reactions. Intensive research has therefore especially been on investigations of biofuel generation and CO2 conversion. In the reductive pathway using dehydrogenase enzymes for the reduction of carbon dioxide, NADH serves as the oxidation equivalent or electron and proton donor. Application of not only such bioinspired but even natural derivatives has become an evolving topic in terms of CO2 reduction and biosynthesis.87 Fig. 8 Three step reaction cascade for the reduction of CO2 to methanol using dehydrogenase enzymes and NADH as the electron and proton donor. Fig. 9 Schematic representation of Lu et al. for the co-immobilization of enzymes in different substrates. Figure reproduced with permission from ref. 104. This journal is © The Royal Society of Chemistry 2017 J. Mater. Chem. A View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Journal of Materials Chemistry A Review Fig. 10 Alginate–silicate based beads containing the three dehydrogenases FateDH, FaldDH and ADH for the reduction of CO2 to methanol with NADH as electron and proton donors. Figure reproduced with permission from ref. 105. 3.1 Reduction reactions with the aid of co-factors Considering CO2 reduction with the aid of dehydrogenase enzymes as biocatalysts, mainly the enzymes formate dehydrogenase, formaldehyde dehydrogenase and alcohol dehydrogenase are reported as they are especially known for the conversion of CO2 via the three step reduction reaction to methanol with the intermediate products formate and formaldehyde. In this reaction cascade, each step is catalyzed by the corresponding enzyme. As a rst step formate dehydrogenase (FateDH) catalyzes the conversion to formic acid or formate. This is followed by reducing formate to formaldehyde by means of formaldehyde dehydrogenase (FaldDH). The last step, the reduction of formaldehyde to methanol occurs with alcohol dehydrogenase (ADH) as the catalyst responsible. Each of these steps is a two electron reduction. In natural processes electrons and required protons are delivered from the oxidation of a sacricial co-enzyme such as nicotinamide adenine dinucleotide (NADH) (Fig. 8).85,94 Table 2 Related to the origin of the enzyme or microorganism that the dehydrogenases are isolated from, the kinetics of these reactions are inuenced to certain extents (eqn (1)). In particular the performance of formate dehydrogenase is highly dependent on its source and this is crucial as it catalyzes the initial step of the reduction of CO2. As it has been found e.g. by Kim and co-workers, formate dehydrogenase obtained from Candida boidinii requires a concentration of the co-factor NADH for efficient activity in a certain range. The activity, however, is inhibited when NADH concentrations were too low or too high. Moreover, they found that formate generation is kinetically favored when CO2 is used directly and not as carbonate. Also the active site of formate dehydrogenase plays a noteworthy role in the kinetics of such reactions. Formate dehydrogenase from Pseudomonas oxalaticus or the tungsten containing FateDH from Syntrophobacter fumaroxidans e.g. are both NADH independent but were found to be very unstable and inactive in the presence of oxygen.95 These dependencies on the active site moreover dene the thermodynamic feasibility of the reduction reactions, as the required potential or chemical energy respectively for the reduction is either dened by NADH or other redox species involved that enable electron transfer.96 Nevertheless, these rather simple reduction reactions using formate, formaldehyde and alcohol dehydrogenase could be imitated in several experimental approaches and delivered promising results. However, what has been found early from the use of enzymes was in general their low chemical stability when exposed to air or kept in inappropriate solvents or pH.97,98 Several studies therefore are focused on stabilizing the enzymes by e.g. encapsulation for better performance. Furthermore, immobilization provides the possibility to reuse the biocatalysts as they can be separated more easily from the reaction solution and products. First approaches of enzyme immobilization were reported by Heichal-Segal et al. using alginate-silicate gel matrices for Summary for co-factor supported enzyme catalyzed CO2 conversion approaches Enzyme Co-factor/mediator Carbonic anhydrase Biotin carboxylase Phosphoenolpyruvate carboxylase Isocitrate dehydrogenase Product Reference HCO3 Carboxyphosphate Oxalacetate Isocitrate Addo et al.,57 Hawkins et al.46 Alissandratos et al.84 Alissandratos et al.84 Alissandratos et al.,84 Sugimura et al.86 Alissandratos et al.,84 Aresta et al.85 Ribulose-1,6-bisphosphatecarboxylase/oxygenase Carbon monoxide dehydrogenase Formate dehydrogenase Mg 3D-Phosphoglycerat Ferrocene NADH Carbon monoxide Formate Formaldehyde dehydrogenase NADH Formaldehyde Alcohol dehydrogenase NADH Methanol J. Mater. Chem. A Tommasi et al.,88 Ruschig et al.,94 Aresta et al.,85 Kim et al.,95 Baskaya et al.,96 Obert et al.,101 Lu et al.,104 Dibenedetto et al.,105 Luo et al.106 Aresta et al.,85 Baskaya et al.96 Obert et al.,101 Xu et al.,104 Dibenedetto et al.,105 Luo et al.106 Aresta et al.,85 Baskaya et al.,96 Obert et al.,101 Xu et al.,103 Dibenedetto et al.,105 Luo et al.106 This journal is © The Royal Society of Chemistry 2017 View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Review glucosidase immobilization to prevent it from thermal and chemical denaturation.99 Later on Aresta et al. investigated enzymes immobilized in agar as well as polyacrylamide, pumice and zeolite materials concerning stability and activity of the carboxylase enzyme to be used for the synthesis of benzoic acid from phenol and CO2. They found that the immobilized enzymes remained active for over one week and they could obtain over 90% conversion efficiency of the phenol derivative to benzoic acid within a few days.100 Among the different immobilization techniques especially encapsulation of enzymes in gels and gel beads is advantageous. Such gels yield a large surface area, provide high chemical stability for the enzymes and reduce costs due to possible reusability of the biocatalysts. Obert and Dave rst presented the immobilization of dehydrogenase enzymes for chemical CO2 reduction in silica sol–gel matrices. They proved the performance of dehydrogenase enzymes for the reduction of CO2 to methanol even when encapsulated in a sol–gel. Moreover an improved yield to up to 90% according to the amount of NADH added was obtained for methanol generation.101 Also sol–gel materials based on phospholipids were found to be viable even for the co-immobilization of all three dehydrogenase enzymes for the conversion of CO2 to methanol.102 Furthermore Lu and co-workers investigated the immobilization of dehydrogenases separately and together in beads of an alginate–silica hybrid gel. They moreover showed the dependence on gel constitution for the chemical CO2 reduction to methanol and formate and improved performance for the hybrid gel (Fig. 9). For silica containing sol–gels an even higher yield was observed. Additionally co-encapsulation of the enzymes in the gel improved the methanol yield further. Fig. 9 shows the encapsulation of enzymes in different systems, screened by the group of Lu.103,104 In the work of Dibenedetto et al. NADH regeneration and the use of new hybrid technologies, by merging biological and chemical technologies, for the conversion of CO2 to methanol with the dehydrogenase cascade, co-immobilized in alginate based beads (Fig. 10), were investigated. They used an alginate– silicate based sol–gel as well for the immobilization.105 In a recent study polymeric sheets were used and the co-immobilization of formate, formaldehyde and alcohol dehydrogenase was presented for the reduction cascade of CO2. Luo and co-workers showed separate as well as co-immobilization in polymeric membranes with efficient and comparable yields for methanol generation.106 Table 2 summarizes all discussed co-factor dependent enzyme catalyzed CO2 conversion approaches. However, in Fig. 11 Reduction mechanisms of CO2 catalyzed by dehydrogenase enzymes. 3-Step reduction of CO2 to methanol, direct electron transfer to the enzyme without any co-factor. This journal is © The Royal Society of Chemistry 2017 Journal of Materials Chemistry A most of these enzymatic CO2 reduction approaches, a co-factor such as NADH was required as the oxidation equivalent and as electron and proton donors for the reduction reactions. This is because basically reactions of such processes follow the Michaelis–Menten equation for enzyme kinetics and are demonstrated in eqn (1). [E] + [S] + [NADH] 4 [E + S + NADH]* 4 [E] + [P] + [NAD+] (1) In this equation E denotes the enzyme, S the substrate (such as CO2 and its intermediate reduction products), and P the nal product.107 As common in enzyme kinetics the use of dehydrogenase enzymes for CO2 reduction with the aid of NADH can be described with the formation of an intermediate state. In this state the substrate and co-enzyme both link to the (metal) active site of the enzyme. This is crucial for the reduction of the substrate due to enabled proton and electron transfer via the active site of the enzyme within the intermediate state. Nevertheless, in those processes NADH is irreversibly converted to its oxidized form NAD+. In experimental approaches co-factors therefore are sacricial and need to be regenerated or steadily supplied to drive reactions repeatedly or even continuously. This irreversible co-factor oxidation limits dehydrogenase applications for CO2 reduction to small scale experiments due to high costs of NADH synthesis and regeneration. Depletion of co-factors and related high costs for regeneration, additional process steps and high amounts of required materials therefore give rise to intensive research for co-factor substitution. Taking the role of co-factors during redox reactions into account, one has to consider especially the electron transfer, which is crucial. A primary goal is therefore to nd a pathway that provides direct electron transfer but does not require a sacricial co-factor. To address this issue electrochemical approaches draw particular attention. Proposed electron transfer from an electrode to formate dehydrogenase during reduction of CO2 to formate, reported by Reda et al. Figure reproduced with permission from ref. 1 and 111. Fig. 12 J. Mater. Chem. A View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Journal of Materials Chemistry A Review Preparation procedure for the immobilization of enzymes on a carbon felt electrode. Encapsulation of dehydrogenases in an alginate– silicate hybrid gel (A). Blank carbon felt electrode (B) that is soaked with the alginate–silicate mixture and precipitated in CaCl2 (C). Alginate covered carbon felt electrode after precipitation (D). Fig. 13 3.2 Electro-enzymatic reduction reactions The implementation of electro-enzymatic processes instead of NADH as the electron donor is of great interest as this offers the possibility of direct electron injection into enzymes without any co-factor or shuttle. Similar to bio-electrochemical applications using microorganisms, electro-enzymatic applications furthermore open the way toward renewable energy storage and greenhouse gas conversion into articial fuels at the same time. Fig. 11 demonstrates the possible reduction route using direct electron injection from an electrode instead of NADH as the electron and proton donor. In the electrochemical process electrons are provided by an external energy source and the cathode and protons are delivered from aqueous electrolyte solutions. An approach to address enzymes electrochemically has been presented e.g. by Minteer and co-workers. They showed direct electron transfer from enzymes for biofuel cells. In a further study they used dehydrogenase enzymes in the opposite direction for electron uptake due to their viability to be electrochemically addressed. They depicted this technique for the bio-electrocatalytic regeneration of NADH supported by (poly)neutral red.57,108 Amao and Shuto demonstrated the application of viologen modied FateDH immobilized on an indium-tin oxide electrode for the reduction of CO2 to formate in an articial photosynthesis approach. Viologen served as the support for the electron transfer between the electrode and enzyme. This approach delivered promising results for a bio-electrocatalytic approach with production rates of up to 7.6 mmol formate/h, when the length of the viologen-containing carbon chain between electrode and enzyme was increased.109 Further, for the direct reduction of CO2 Kuwabata et al. presented an electrochemical approach for the conversion to methanol using dehydrogenase enzymes and methylviologen as well as pyrroloquinolinequinone as supporting electron mediators. Also in this work the viologen derivative turned out to be highly favorable as an electron mediator with current efficiencies up to 91%.110 Reda et al. later demonstrated the electrochemical CO2 reduction in a heterogeneously catalyzed approach, using tungsten containing formate dehydrogenase adsorbed on glassy J. Mater. Chem. A carbon without any electron shuttle. This work yielded faradaic efficiencies of 97% and higher. They moreover proposed a mechanism for the electron transfer to the enzyme and subsequent reduction of CO2 as depicted in Fig. 12.111 Also Bassegoda et al. later used formate dehydrogenase with a metal active site for the reversible conversion of formate and CO2. They suggest the molybdenum containing FateDH as a highly active electrocatalyst for such redox processes.112 We recently investigated the immobilization of dehydrogenases encapsulated in an alginate based matrix on a carbon felt electrode (Fig. 13). We found that for the immobilization of a single enzyme, namely alcohol dehydrogenase, conversion of an aldehyde to the corresponding alcohol is possible simply due to direct electron injection without any co-factor or electron shuttle.113 In a similar approach, using this technique for electrode preparation, we then investigated the co-immobilization of all three dehydrogenases on such electrodes for the conversion of CO2 to methanol (Fig. 14). We obtained faradaic yields of 10% for methanol generation from electro-enzymatic CO2 reduction.114 Besides electrochemical methods one also has to mention here the possibility of photo-reduction of CO2. For this charges Fig. 14 Scheme of the setup of an electrochemical cell for electroenzymatic CO2 reduction. Electrons are injected directly into the enzymes, which are encapsulated in an alginate–silicate hybrid gel (green) and immobilized on a carbon felt working electrode. CO2 is reduced to methanol at the working electrode. Oxidation reactions take place at the counter electrode.114 This journal is © The Royal Society of Chemistry 2017 View Article Online Review Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Table 3 Journal of Materials Chemistry A Summary of electrochemical approaches using enzymes for the reduction of CO2 via direct electron injection Enzyme Immobilization Co-factor/mediator Product Formate, formaldehyde, alcohol dehydrogenase Formate dehydrogenase No Methanol Glassy carbon Yes Formate dehydrogenase, No methanol dehydrogenase Formate dehydrogenase Yes Formate dehydrogenase Alcohol dehydrogenase Yes Yes Formate, formaldehyde, alcohol dehydrogenase Yes NADH/neutral red Electrode Cathodic potential 0.8 V vs. SCE (0.556 V vs. SHE) Viologen Formate ITO 0.55 V vs. Ag/AgCl (0.34 V vs. SHE) Methylviologen, Methanol Glassy carbon 0.7–0.9 V vs. SCE pyrolloquinolinequinone (0.456–0.656 vs. SHE) None Formate Glassy carbon 0.41–0.81 V vs. Ag/AgCl (0.2–0.6 V vs. SHE) None Formate Graphite-epoxy 0.5–0.6 V vs. SHE None Butanol Carbon felt 0.6 V vs. Ag/AgCl (0.39 V vs. SHE) None Methanol Carbon felt 1.2 V vs. Ag/AgCl (0.99 V vs. SHE) or electrons are provided by a light source. Upon light excitation electrons can be transferred from a semiconductor electrode to an enzyme. Such an approach has been presented by the group of Chaudhary et al. who used CdS nanoparticles assembled with carbon monoxide dehydrogenase. They obtained turn over frequencies for the conversion of CO2 to CO up to 22 000 aer 5 hours. This is higher than what was observed for synthetic catalysts but still smaller than the maximum capability of enzymes.115 Such approaches as the electrochemical or photochemical reduction of CO2, utilizing electrodes with immobilized dehydrogenase enzymes and without any co-enzyme or mediator required, depict a sustainable route towards biocompatible reduction of the greenhouse gas CO2. This approach further shows a viable and promising strategy for the substitution of expensive co-factors. This makes enzymatic redox reactions also attractive for industrial applications not only in the elds of CO2 reduction and renewable energy storage but also for extended applications like the production of food additives and base chemicals catalyzed by different enzymes. A comparison of the different electro-enzymatic approaches is displayed in Table 3. 4. Conclusions Biocatalysis and especially bio-electrocatalysis are interesting and promising methods for CO2 recycling into chemicals and articial fuels. Existing industrial methods for this purpose and chemically synthesized catalysts oen lack in selectivity or yield for the obtained products. Furthermore, high temperatures or pressures may be required in such processes. In contrast, biocatalysts usually work under mild conditions at room temperature and atmospheric pressure. As a result, direct application of biocatalysts has become more and more industrially attractive. Utilization of microorganisms or enzymes therefore paves the way for CO2 conversion and generation of various chemicals and biofuels with high yield and high selectivity. As presented in various studies, biocatalytic especially bio-electrocatalytic approaches offer high economic potential. Even though there is still intensive work required for optimization to possible large scale application, small scale approaches already showed that This journal is © The Royal Society of Chemistry 2017 Reference Minteer et al.57 Amao et al.109 Kuwabata et al.110 Reda et al.111 Bassegoda et al.112 Schlager et al.113 Schlager et al.114 not only high yields can be obtained, but even costs can be reduced for electrochemical approaches without co-factors or hydrogen equivalents like NADH or H2. In particular in the case of the co-factor NADH cost reduction for CO2 conversion can be obtained by several orders of magnitude when electrolysis approaches were used instead. Such useful recycling of CO2 moreover displays a viable approach for large scale chemical storage of renewable energies, as these can serve as a source for the electrical energy supply. Acknowledgements The authors acknowledge nancial support by the Austrian Science Foundation (FWF) within the Wittgenstein Prize (Z222-N19), FFG within the project CO2TRANSFER (848862) and REGSTORE project which was funded under the EU Programme Regional Competitiveness 2007–2013 (Regio 13) with funds from the European Regional Development Fund and by the Upper Austrian Government. Notes and references 1 C. D. Keeling, Geochim. Cosmochim. Acta, 1958, 13, 322–334. 2 Scripps Institution of Oceanography, UC San Diego, 2016, https://scripps.ucsd.edu/programs/keelingcurve/, August 30th, 2016. 3 H. Craig, Tellus, 1957, 9, 1–17. 4 B. Weinstock, Science, 1969, 166, 224–225. 5 M. Aresta and G. Forti, Carbon Dioxide as a Source of Carbon: Biochemical and Chemical Uses, D. Reidel Publishing Company, Dordrecht, 1987. 6 M. Aresta, Carbon Dioxide as Chemical Feedstock, WileyVCH, Weinheim, 2010. 7 G. A. Olah, A. Goeppert and G. K. Surya Prakash, Beyond Oil and Gas: The Methanol Economy, Wiley-VCH, Weinheim, 2nd edn, 2009. 8 M. Aresta, Carbon Dioxide Recovery and Utilization, Kluwer Academic Publishers, Dordrecht, 2003. 9 J. Qiao, Y. Liu, F. Hong and J. Zhang, Chem. Soc. Rev., 2014, 43, 631–675. J. Mater. Chem. A View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Journal of Materials Chemistry A 10 S. Freguia, B. Virdis, F. Harnisch and J. Keller, Electrochim. Acta, 2012, 82, 165–174. 11 J. G. Zeikus, Annu. Rev. Microbiol., 1980, 34, 423–464. 12 J. R. Philips, E. C. Clausen and J. L. Gaddy, Appl. Biochem. Biotechnol., 1994, 45/46, 145–157. 13 R. Kerby and J. G. Zeikus, Curr. Microbiol., 1983, 8, 27–30. 14 H. L. Drake, A. S. Gößner and S. L. Daniel, Ann. N. Y. Acad. Sci., 2008, 1125, 100–128. 15 S. L. Lee, H. Chou, T. S. Ham, T. S. Lee and J. D. Keasling, Curr. Opin. Biotechnol., 2008, 19, 556–563. 16 M. Mohammadi, G. F. Najafpour, H. Younesi, P. Lahijani, M. H. Uzir and A. R. Mohamed, Renewable Sustainable Energy Rev., 2011, 15, 4255–4273. 17 M. Mohammadi, H. Younesi, G. Najapour and A. R. Mohamed, J. Chem. Technol. Biotechnol., 2012, 87, 837–843. 18 H. G. Wood, FASEB J., 1991, 5, 156–163. 19 J. S.-C. Liou, D. L. Balkwill, G. R. Drake and R. S. Tanner, Int. J. Syst. Evol. Microbiol., 2005, 55, 2085–2091. 20 I. A. Berg, D. Kockelkorn, W. H. Ramos-Vera, R. F. Say, J. Zarzycki, M. Hügler, B. E. Alber and G. Fuchs, Nat. Rev. Microbiol., 2010, 8, 447–460. 21 H. Younesi, G. Najafpour and A. R. Mohamed, Biochem. Eng. J., 2005, 27, 110–119. 22 S. W. Ragsdale and E. Pierce, Biochim. Biophys. Acta, Gen. Subj., 2009, 1784, 1873–1898. 23 S. W. Ragsdale, Ann. N. Y. Acad. Sci., 2008, 1125, 129–135. 24 B. P. Tracy, S. W. Jones, A. G. Fast, D. C. Indurthi and E. T. Papoutsakis, Curr. Opin. Biotechnol., 2012, 23, 364–381. 25 B. Schiel-Bengelsdorf and P. Dürre, FEBS Lett., 2012, 586, 2191–2198. 26 P. C. Munasinghe and S. K. Khanal, Bioresour. Technol., 2010, 101, 5013–5022. 27 P. Hu, H. Rismani-Yazdi and G. Stephanopoulos, AIChE J., 2013, 7, 405–410. 28 D. K. Kundiyana, M. R. Wilkins, P. Maddipati and R. L. Huhnke, Bioresour. Technol., 2011, 102, 5794–5799. 29 K. Liu, H. K. Atiyeh, B. S. Stevenson, R. S. Tanner, M. R. Wilkins and R. L. Huhnke, Bioresour. Technol., 2014, 151, 69–77. 30 L. Gustavsson and A. Karlsson, Mitig. Adapt. Strateg. Glob. Change, 2006, 11, 935–959. 31 J. Sheehan, Nat. Biotechnol., 2009, 27, 1128–1129. 32 B. E. Rittmann, Biotechnol. Bioeng., 2008, 100, 203–212. 33 U. Deppenmeier, V. Müller and G. Gottschalk, Arch. Microbiol., 1996, 165, 149–163. 34 D. C. Ducat and P. A. Silver, Curr. Opin. Chem. Biol., 2012, 16, 337–344. 35 J. L. Garcia, B. K. C. Patel and B. Ollivier, Anaerobe, 2000, 6, 205–226. 36 M. R. Connor and S. Atsumi, J. Biomed. Biotechnol., 2010, 2010, 1–9. 37 H. Li, P. G. Opgenorth, D. G. Wernick, S. Rogers, T.-Y. Wu, W. Higashide, P. Malati, Y.-X. Huo, K. M. Cho and J. C. Liao, Science, 2012, 335, 1596. 38 M. Straub, M. Demler, D. Weuster-Botz and P. Dürre, J. Biotechnol., 2014, 178, 67–72. J. Mater. Chem. A Review 39 J. Daniell, M. Köpke and S. D. Simpson, Energies, 2012, 5, 5372–5417. 40 E. Marcellin, J. B. Behrendorff, S. Nagaraju, S. DeTissera, S. Segovia, R. W. Palfreyman, J. Daniell, C. LiconaCassani, L. Quek, R. Speight, M. P. Hodson, S. D. Simpson, W. P. Mitchell, M. Köpke and L. K. Nielsen, Green Chem., 2016, 18, 3020–3028. 41 D. K. Kundiyana, R. L. Huhnke and M. R. Wilkins, J. Biosci. Bioeng., 2010, 109, 492–498. 42 J. Song, Y. Kim, M. Lim, H. Lee, J. I. Lee and W. Shin, ChemSusChem, 2011, 4, 587–590. 43 D. R. Lovley and K. P. Nevin, Curr. Opin. Biotechnol., 2013, 24, 385–390. 44 H. Wang and Z. Ren, Biotechnol. Adv., 2013, 31, 1796–1807. 45 K. D. Ramachandriya, D. K. Kundiyana, M. R. Wilikins, J. B. Terrill, H. K. Atiyeh and R. L. Huhnke, Appl. Energy, 2013, 112, 289–299. 46 A. S. Hawkins, P. M. McTernan, H. Lian, R. M. Kelly and M. W. W. Adams, Curr. Opin. Biotechnol., 2013, 24, 376–384. 47 S. Patil, K. Hasan, D. Leech, C. Hägerhäll and L. Gorton, Chem. Commun., 2012, 48, 10183–10185. 48 B. E. Logan, Microbial Fuel Cells, Wiley, New Jersey, 2008. 49 H. L. Ehrlich, Geobiology, 2008, 6, 220–224. 50 M. Rosenbaum, F. Aulenta, M. Villano and T. Angenent, Bioresour. Technol., 2011, 102, 324–333. 51 O. Choi and B.-I. Sang, Biotechnol. Biofuels, 2016, 9, 1–14. 52 D. E. Holmes, T. Mester, R. A. O'Neil, L. A. Perpetua, M. J. Larrahondo, R. Glaven, M. L. Sharma, J. E. Ward, K. P. Nevin and D. R. Lovley, Microbiology, 2008, 154, 1422–1435. 53 T. Yamanaka and Y. Fukumori, FEMS Microbiol. Rev., 1995, 17, 401–413. 54 F. Aulenta, A. Catervi, M. Majone, S. Panero, P. Reale and S. Rossetti, Environ. Sci. Technol., 2007, 41, 2554–2559. 55 K. J. J. Steinbusch, H. V. M. Hamelers, J. D. Schaap, C. Kampman and C. J. N. Buisman, Environ. Sci. Technol., 2010, 44, 513–517. 56 S. Kuwabata, R. Tsuda and H. Yoneyama, J. Am. Chem. Soc., 1994, 116, 5437–5443. 57 P. K. Addo, R. L. Arechederra, A. Waheed, J. D. Shoemaker, W. S. Sly and S. D. Minteer, Electrochem. Solid-State Lett., 2011, 14, E9–E13. 58 H. M. Jensen, A. E. Albers, K. R. Malley, Y. Y. Londer, B. E. Cohen, B. A. Helms, P. Weigele, J. T. Groves and C. M. Ajo-Franklin, PNAS, 2010, 107, 19213–19218. 59 C. M. Ajo-Franklin and A. Noy, Adv. Mater., 2015, 27, 5795–5804. 60 S. Cheng and B. E. Logan, PNAS, 2007, 104, 18871–18873. 61 B. E. Logan, D. Call, S. Cheng, H. V. M. Hamelers, T. H. J. A. Sleutels, A. W. Jeremiasse and R. A. Rozendal, Environ. Sci. Technol., 2008, 42, 8630–8640. 62 S. A. Cheng, D. F. Xing, D. F. Call and B. E. Logan, Environ. Sci. Technol., 2009, 43, 3953–3958. 63 P. Clauwaert, R. Toledo, D. van der Ha, R. Crab, W. Verstraete, H. Hu, K. M. Udert and K. Rabaey, Water Sci. Technol., 2008, 57, 575–579. 64 K. Rabaey and R. A. Rozendal, Nat. Rev. Microbiol., 2010, 8, 706–716. This journal is © The Royal Society of Chemistry 2017 View Article Online Published on 02 December 2016. Downloaded by Johannes Kepler Universitat Linz on 02/02/2017 09:09:45. Review 65 M. Villano, F. Aulenta, C. Ciucci, T. Ferri, A. Giuliano and M. Majone, Bioresour. Technol., 2010, 101, 3085–3090. 66 M. Villano, G. Monaco, F. Aulenta and M. Majone, J. Power Sources, 2011, 196, 9467–9472. 67 Y. Jiang, M. Su, Y. Zhang, G. Zhan, Y. Tao and D. Li, Int. J. Hydrogen Energy, 2012, 38, 3497–3502. 68 L. Jourdin, T. Grieger, J. Monetti, V. Flexer, S. Freguia, Y. Lu, J. Chen, M. Romano, G. G. Wallace and J. Keller, Environ. Sci. Technol., 2015, 49, 13566–13574. 69 G. Zhen, T. Kobayashi, X. Lu and K. Xu, Bioresour. Technol., 2015, 186, 141–148. 70 G. Zhen, X. Lu, T. Kobayashi, G. Kumar and K. Xu, Chem. Eng. J., 2016, 284, 1146–1155. 71 M. C. A. A. Van Eerten-Jansen, A. Ter Heijne, C. J. N. Buisman and H. V. M. Hamelers, Int. J. Energy Res., 2012, 36, 809–819. 72 S. Bajracharya, A. ter Heijne, X. D. Benetton, K. Vanbroekhoven, C. J. N. Buisman, D. P. B. T. B Strik and D. Pant, Bioresour. Technol., 2015, 195, 14–24. 73 S. Schlager, M. Haberbauer, A. Fuchsbauer, C. Hemmelmair, L. M. Dumitru, G. Hinterberger, H. Neugebauer and N. S. Saricici, ChemSusChem, 2016, 9, DOI: 10.1002/cssc.201600963. 74 S. E. Lowe, M. K. Jain and G. Zeikus, Microbiol. Mol. Biol. Rev., 1993, 57, 451–509. 75 H. A. Hansen, J. B. Varley, A. A. Peterson and J. K. Norskov, J. Phys. Chem. Lett., 2013, 4, 388–392. 76 B. Mondal, J. Song, F. Neese and S. Ye, Curr. Opin. Chem. Biol., 2015, 25, 103–109. 77 E. E. Benson, C. P. Kubiak, A. J. Sathrum and J. M Smieja, Chem. Soc. Rev., 2008, 38, 89–99. 78 M. Aresta and A. Dibenedetto, Rev. Mol. Biotechnol., 2002, 90, 113–128. 79 M. Aresta and A. Dibenedetto, Dalton Trans., 2007, 2975–2992. 80 V. C.-C. Wang, S. W. Ragsdale and F. A. Armstrong, ChemBioChem, 2013, 14, 1845–1851. 81 M. Rakowski Dubois and D. L. Dubois, Acc. Chem. Res., 2009, 42, 1974–1982. 82 H. R. Mahler and J. Douglas, J. Am. Chem. Soc., 1957, 79, 1159–1166. 83 Ö. Almarsson and T. C. Bruice, J. Am. Chem. Soc., 1993, 115, 2125–2138. 84 A. Alissandratos and C. J. Easton, Beilstein J. Org. Chem., 2015, 11, 2370–2387. 85 M. Aresta, A. Dibenedetto and C. Pastore, Environ. Chem. Lett., 2005, 3, 113–117. 86 K. Sugimura, S. Kuwabata and H. Yoneyama, J. Am. Chem. Soc., 1989, 111, 2361–2362. 87 J. Shi, Y. Jiang, Z. Jiang, X. Wang, X. Wang, S. Zhang, P. Han and C. Yang, Chem. Soc. Rev., 2015, 44, 5981–6000. 88 I. Tommasi, M. Aresta, P. Giannoccaro, E. Quaranta and C. Fragale, Inorg. Chim. Acta, 1998, 272, 38–42. 89 W. Shin, S. H. Lee, J. W. Shin, S. P. Lee and Y. Kim, J. Am. Chem. Soc., 2003, 125, 14688–14689. 90 C. Mateo, J. M. Palomo, G. Fernandez-Lorente, J. M. Guisan and R. Fernandez-Lafuente, Enzyme Microb. Technol., 2007, 40, 1451–1463. This journal is © The Royal Society of Chemistry 2017 Journal of Materials Chemistry A 91 G. T. R. Palmore, H. Bertschy, S. H. Bergens and G. M. Whitesides, J. Electroanal. Chem., 1998, 443, 155–161. 92 Z. Shaked and G. M. Whitesides, J. Am. Chem. Soc., 1980, 102, 7104–7105. 93 W. Neuhauser, M. Steininger, D. Haltrich, K. D. Kulbe and B. Nidetzky, Biotechnol. Bioeng., 1998, 60, 277–282. 94 U. Ruschig, U. Müller, P. Willnow and T. Höpner, Eur. J. Biochem., 1976, 70, 325–330. 95 S. Kim, M. K. Kim, S. H. Lee, S. Yoon and K.-D. Jung, J. Mol. Catal. B: Enzym., 2014, 102, 9–15. 96 F. S. Baskaya, X. Zhao, M. C. Flickinger and P. Wang, Appl. Biochem. Biotechnol., 2010, 162, 391–398. 97 L. Betancor and H. R. Luckari, Trends Biotechnol., 2008, 26, 566–572. 98 P. V. Iyer and L. Ananthanaryan, Process Biochem., 2008, 43, 1019–1032. 99 O. Heichal-Segal, S. Rappoport and S. Braun, Bio/ Technology, 1995, 13, 798–800. 100 M. Aresta, E. Quaranta, R. Liberio, C. Dileo and I. Tommasi, Tetrahedron, 1998, 54, 8841–8846. 101 R. Obert and B. C. Dave, J. Am. Chem. Soc., 1999, 121, 12192– 12193. 102 R. Cazelles, J. Drone, F. Fajula, O. Ersen, S. Moldovan and A. Galarneau, New J. Chem., 2013, 37, 3721–3730. 103 S.-W. Xu, Y. Lu, J. Li, Z.-y. Jiang and H. Wu, Ind. Eng. Chem. Res., 2006, 45, 4567–4573. 104 Y. Lu, Z.-Y. Jiang, S.-W. Xu and H. Wu, Catal. Today, 2006, 115, 263–268. 105 A. Dibenedetto, P. Stufano, W. Macyk, T. Baran, C. Fragale, M. Costa and M. Aresta, ChemSusChem, 2012, 5, 373–378. 106 J. Luo, A. S. Meyer, R. V. Matieu and M. Pinelo, New Biotechnol., 2015, 32, 319–327. 107 L. Michaelis and M. L. Menten, Biochem. Z., 1913, 49, 333– 369. 108 S. D. Minteer, B. Y. Liaw and M. J. Cooney, Curr. Opin. Biotechnol., 2007, 18, 228–234. 109 Y. Amao and N. Shuto, Res. Chem. Intermed., 2013, 40, 3267– 3276. 110 S. Kuwabata, R. Tsuda and H. Yoneyama, J. Chem. Soc., 1994, 116, 5437–5443. 111 T. Reda, C. M. Plugge, N. J. Abram and J. Hirst, PNAS, 2008, 105, 10654–10658. 112 A. Bassegoda, C. Madden, D. W. Wakerley, E. Reisner and J. Hirst, J. Am. Chem. Soc., 2014, 136, 15473–15476. 113 S. Schlager, H. Neugebauer, M. Haberbauer, G. Hinterberger and N. S. Saricici, ChemCatChem, 2015, 7, 967–971. 114 S. Schlager, L. M. Dumitru, M. Haberbauer, A. Fuchsbauer, H. Neugebauer, D. Hiemetsberger, A. Wagner, E. Portenkirchner and N. S. Saricici, ChemSusChem, 2016, 9, 631–635. 115 Y. S. Chaudhary, T. W. Woolerton, C. S. Allen, J. H. Warner, E. Pierce, S. W. Ragsdale and F. A. Armstrong, Chem. Commun., 2012, 48, 58–60. J. Mater. Chem. A
© Copyright 2025 Paperzz