Next Article in Journal
Comparison of Homemade and Commercial Plant-Based Drinks (Almond, Oat, Soy) Fermented with Yogurt Starter Culture for Fresh Consumption
Next Article in Special Issue
Mixotrophic Syngas Conversion Enables the Production of meso-2,3-butanediol with Clostridium autoethanogenum
Previous Article in Journal
Advances in Droplet-Based Microfluidic High-Throughput Screening of Engineered Strains and Enzymes Based on Ultraviolet, Visible, and Fluorescent Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microbial Electrosynthesis Using 3D Bioprinting of Sporomusa ovata on Copper, Stainless-Steel, and Titanium Cathodes for CO2 Reduction

Biochemical Process Engineering, Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, 971-87 Luleå, Sweden
*
Authors to whom correspondence should be addressed.
Fermentation 2024, 10(1), 34; https://doi.org/10.3390/fermentation10010034
Submission received: 30 November 2023 / Revised: 23 December 2023 / Accepted: 29 December 2023 / Published: 30 December 2023
(This article belongs to the Special Issue Microbial Fixation of CO2 to Fuels and Chemicals)

Abstract

:
Acetate can be produced from carbon dioxide (CO2) and electricity using bacteria at the cathode of microbial electrosynthesis (MES). This process relies on electrolytically-produced hydrogen (H2). However, the low solubility of H2 can limit the process. Using metal cathodes to generate H2 at a high rate can improve MES. Immobilizing bacteria on the metal cathode can further proliferate the H2 availability to the bacteria. In this study, we investigated the performances of 3D bioprinting of Sporomusa ovata on three metal meshes—copper (Cu), stainless steel (SS), and titanium (Ti), when used individually as a cathode in MES. Bacterial cells were immobilized on the metal using a 3D bioprinter with alginate hydrogel ink. The bioprinted Ti mesh exhibited higher acetate production (53 ± 19 g/m2/d) at −0.8 V vs. Ag/AgCl as compared to other metal cathodes. More than 9 g/L of acetate was achieved with bioprinted Ti, and the least amount was obtained with bioprinted Cu. Although all three metals are known for catalyzing H2 evolution, the lower biocompatibility and chemical stability of Cu hampered its performance. Stable and biocompatible Ti supported the bioprinted S. ovata effectively. Bioprinting of synthetic biofilm on H2-evolving metal cathodes can provide high-performing and robust biocathodes for further application of MES.

1. Introduction

There have been continuous global attempts to reduce carbon dioxide (CO2) emissions. CO2 can be captured and transformed chemically or biologically into useful compounds. Microorganisms are being utilized for the electrocatalytic conversion of CO2 to compounds like methane (CH4), acetate, and ethanol at the cathode with the input of electric power. Bacterial genera such as Sporomusa and Clostridium have been shown to use electric current to turn CO2 into acetate and other multi-carbon compounds at the cathode in a system called microbial electrosynthesis (MES) [1,2]. MES has gained high interest in recent years for capturing and utilizing carbon in valuable chemicals or fuels [3]. As such, it shows the potential to reduce our reliance on fossil fuels and combat climate change [4]. However, current MES technology is not yet economically viable. Only bulk products such as acetate and/or short-chain carboxylates are reported as being produced in MES [5]. Improving the performance of MES systems will be crucial in making these technologies more cost-effective.
The productivity of MES reactors has been successfully increased using a variety of methods, such as S. ovata adaptation to methanol-containing growth medium [6] and structural optimization of the microbial and cathode interface [7,8,9]. Studies have focused on the utilization of mixed cultures derived from natural anaerobic sources for MES [10,11], and long-term enrichment of the microbiome and biofilm has led to significant performance gains [12,13,14]. Studies obtaining the highest productivity have focused on modifying the cathode in various ways. There have been two main approaches. The first is to increase the biomass concentration on the cathode [13,15,16,17,18]. The second approach is to modify the electrochemical properties of the cathode, either through the surface modification of carbonaceous electrodes or through the addition of metal particles or using metallic electrodes [12,19,20,21]. One of the difficulties in designing ideal MES cathodes is that they need to meet several requirements simultaneously, namely high electrical conductivity, good biocompatibility, and a high surface area, while having a high durability and low production cost [22]. Some of these properties tend to be difficult to meet; for example, copper foam electrodes fulfill all the requirements, except for biocompatibility, since copper can be toxic to several microorganisms [23]. But investigations on suitable and cost-effective metal cathodes for MES are still lacking. S. ovata has often been used as a model organism in MES, and is well known for its use in microbial electrosynthesis reactors, producing mainly acetate [1,24]. Mixed cultures might be more feasible for large-scale operations, but the large variability of microbial structure and function in mixed cultures would add uncertainties in the performance of cathode materials [25]. S. ovata is, however, not capable of forming the dense biofilms as observed in other electroactive bacteria such as Geobacter sulfurreducens [26,27]. That being said, increasing the amount of microbial cells directly on the electrode has shown significant increase in production; for example, the use of silicon nanowire arrays to entrap the cells [18], electrode modification to make it more biocompatible [28], and also the use of 3D-bioprinted biofilms [16] are the ways to increase cell density in cathode biofilm. For 3D-bioprinting biofilms, a concentrated cell suspension is mixed with a cross-linkable hydrogel and printed directly on the cathode, forming a dense biofilm in approximately a third of the normal time required (less than two days, compared to a week typically required) [16,29]. Bioprinted biofilms, for example, allow for significant biofilm growth, even when using microbes that do not naturally form thick biofilms, such as S. ovata, and has been shown to greatly increase acetate production in MES with an average acetate production of 47 ± 5.1 g/m2/d (0.31 ± 0.55 g/L/d) [16].
Electron uptake from the cathode in MES systems occurs in two ways, direct electron uptake by the biocatalyst or indirect electron uptake via H2 production [30]. Although the core mechanics of electron uptake are still uncertain, it has become widely accepted that indirect electron uptake is the main driving force for MES [24]. MES reactors have been shown to enhance the total production of hydrogen, even when merely using spent media, and it has been proposed to occur via the aid of secreted hydrogenases or catalytically active precipitates on the electrodes [31]. This is most likely the reason for the increased MES productivity while enhancing the biofilm thickness. The gas–liquid mass-transfer is also seen as the limiting factor in the uptake of H2 and CO2, with most of the gas escaping the reactor, especially when considering the low solubility of H2. Several reactor designs have been tried to improve the gas to liquid mass-transfer [13,20,21]. This makes the combination of H2 evolving electrodes and well-established biofilms especially productive, since the H2 can be consumed before it can escape the reactor. Metallic cathodes are generally considered the most effective for H2 evolution via electrochemical water splitting, and non-precious HER catalysts such as Ti, Cu, and stainless steel (SS) can be used for MES reactors [32]. It was evident from Bajracharya et al. [21] and Krige et al. [16] that hydrogen is required for the CO2 reduction, and the rates were high when hydrogen-evolving Ti mesh was introduced as a second cathode in MES [21]. Higher rates of acetate production were reported by using Sporomusa ovata in 3D bioprinting [16]. Based on the results from these studies, we are proposing to use bioprinting of bacterial cells directly on the Cu, SS, and Ti metal cathodes for achieving a high rate of hydrogen evolution. However, there are some concerns that biofilm formation can be hindered by antimicrobial metal ion formation through the oxidation of such cathodes during the long-term operation of MES [33,34]. Few studies have used metallic cathodes for MES, but to our knowledge no studies have been reported using 3D-bioprinted biofilms with metallic electrodes. By combining synthetic 3D-printed S. ovata biofilms and H2-evolving metallic cathodes, in this study we are investigating the MES performances of different metal cathodes to minimize the anti-microbial effects, while still maintaining a thick biofilm on the cathode. In addition, we also recirculated the headspace gas to the catholyte in all reactors, to better reutilize the escaping CO2 and H2 gases.

2. Materials and Methods

2.1. Cultivation and Enrichment of S. ovata

Sporomusa ovata cultures (DSM 2662) were purchased from Deutsche Sammlung Mikroorganismen und Zellkulturen (DSMZ), Braunschweig, Germany, and cultivated anaerobically in standard DSM 311 media without casitone. The subsequent cultures were carried out under a mixture of H2, CO2, and N2 (60:20:20) in the same media without casitone and betaine. To stimulate the cells for autotrophic growth, only CO2 and HCO3 were used as a carbon source and H2 was used as an electron donor. A small amount of yeast extract (0.5 g/L) was added to the media to help with initial biomass growth.

2.2. MES Reactor Construction and Operation

All experiments were carried out in standard H-type, dual-chambered glass reactors (270 mL per chamber), separated by a Nafion 117 ion exchange membrane (Ion Power GmbH, München-Flughafen, Germany). DSMZ 311 media without organic substrates were used as electrolytes in MES. The reactors were stirred using a magnetic stirrer and kept in a temperature-controlled plexiglass box (35 °C). Graphite rods (Schunk Carbon Technology AB, Lenhovda, Sweden; diameter: 20 mm, height: 70 mm) were used as an anode. The cathode chamber was equipped with an Ag/AgCl reference electrode (eDAQ, Denistone, East NSW, Australia). The cathodes consisted of 25 mm × 25 mm sized square-shaped metal meshes of copper and stainless steel. The titanium mesh was of size 28 mm × 70 mm. Final productivity is therefore standardized with regards to area and reported as area-based productivity to account for differences in electrode area. Each metal mesh cathode had a wire of the same material as a current collector, which also held the cathode by being pierced through the butyl stopper. The copper electrode was first annealed at 300 °C for 8 h to form a CuOx layer [35]. Annealed electrodes are efficient in catalyzing CO2 reduction and are less sensitive to the deactivation phenomena that destroy bulk metal electrodes [23,35].
A gas bag (~3L volume) filled with 100% CO2 was added to the cathode chamber, and the gas was recirculated to the catholyte using a peristaltic pump. The gas bag had two opening ports, one for collecting the cathode headspace gas from the reactor and another one for pumping the gas back to the catholyte. The recirculation of gas from the gas bag to the catholyte was carried out at 6–10 mL/min using a peristaltic pump. With the operation of the MES, hydrogen gas was produced and accumulated in the gas bag, thus, H2 and CO2 gas mixture was being collected in the gas bag and recirculated. The reactors were connected to a multi-channel potentiostat (MultiEmStat+, Palmsens, Houten, The Netherlands). The metal cathodes with 3D-bioprinted S. ovata were used as working electrodes which were poised at −0.8 vs. Ag/AgCl or more negative potentials to ensure cathodic H2 evolution. At the start-up, the cathode chamber was first sparged with 100% CO2 for 5 min. The composition of gas in the gas bag changed throughout the MES run, since excess H2 was produced, and the CO2 was converted to acetate by bacteria. In addition, the volume of gas in the bag also decreased during operation of MES and the bags were subsequently refilled with 100% CO2 every 3–4 days to maintain sufficient CO2 availability. The anode side was not sparged with any gas, but was in continuous stirring using a stirring magnet. During the MES operation, O2 evolution was occurring at the anode. There was a Nafion 117 proton exchange membrane (PEM) separating the anode and cathode to prevent the passing of O2 to the cathode side.

2.3. Bio-Printing of Synthetic Biofilms

A synthetic biofilm directly onto the cathode was created on the metal-mesh cathodes by 3D-bioprinting an alginate-based hydrogel, containing cellulose nanofibrils (Cellink bioink, Cellink, Sweden), inoculated with S. ovata. The hydrogel preparation and printing were performed in an anaerobic hood, to maintain anaerobic conditions as previously described [16]. A concentrated solution of S. ovata was prepared by centrifugation of the entire volume of a 100 mL inoculum culture down to a pellet and resuspending the pellet in 0.3 mL media. The 0.3 mL of cell suspension was then mixed with 3 mL of gel. The gel was then used to print several electrodes, to maintain similar starting cell concentrations between the replicates. The cathodes were held in place by a sterilized 3D-printed fixture, and a single layer of hydrogel (15 mm × 15 mm, ~0.5 mm thick) was printed onto the cathode using the 3D printer. An ionic crosslinking agent (CELLINK, Gothenburg, Sweden) was then applied over the printed layer to cross-link the gel before fixing the cathode in the reactor, containing 270 mL of DSMZ 311 media as catholyte. A photograph of the bioprint layer on the Cu cathode is shown in Figure 1.

2.4. Chemical Analyses

Samples of catholyte (1~2 mL) were collected from the cathode chambers and filtered through 0.2 µm syringe filters. Acetate and/or other volatile fatty acids were analyzed using a high-performance liquid chromatograph (HPLC) (PerkinElmer Inc., Waltham, MA, USA), equipped with an Aminex HPX-87H column (Bio-Rad Laboratories Inc., Hercules, CA, USA), and a refractive index detector. A solution of 5 mM H2SO4 (0.6 mL/min) was used as the mobile phase and the column was kept at 65 °C.

2.5. Calculations

The performance of the MES was evaluated in terms of rate of production of the value-added products and cathodic columbic efficiency (CE). In this study, only acetate production was considered for calculations. The rate of production of any reduced compound (P) in g/L/d in a batch process is given by:
Pi,t = (Ci, t − Ci, t-1)/Δt
where subscripts i refer to any product from CO2 reduction and subscripts t and t-1 refer to subsequent samples. C is concentration (g/L) of product (here acetate), Δt is the time difference (days) between the two samples.
Number of moles of acetate produced was calculated as:
nacetate,t =Vcat × (Cacetate,t − Cacetate, t-1)/Macetate
where n is the number of moles of a product, Vcat is the total volume of the catholyte (L), and M is its molecular weight (g/mol). Electric equivalents of acetate (in moles) were calculated from the electric current as:
C u r r e n t   e q u i v a l e n t   a c e t a t e = 0 t I d t 8 F
where I is the current (in ampere), 8 is the moles of electrons needed to produce one mole of acetate, F is Faraday’s constant F = 96,485 C/mol e−. The current in equivalent acetate was later converted to g/L to match with the measured acetate concentration profile.
Coulombic efficiency (CE) accounts specifically for the fraction of electrons that ended up as acetate produced from CO2 reduction. CE for CO2 reduction is the efficiency of capturing the electron from the electric currents to the product(s).
CE = nacetate/Current equivalent acetate × 100%

2.6. Scanning Electron Microscopy (SEM)

Small pieces of metal cathodes, with the bioprint, were cut off, and the biofilm was then fixed by submerging it in 2.5% w/v glutaraldehyde overnight and then freeze-dried to prepare for SEM imaging. Biofilms on the metal cathodes were visualized in FEI Magellan 400 SEM, to study the structure and morphologies.

3. Results

After centrifugation, the cells of S. ovata were bioprinted on various metal cathodes instead of carbon-based electrodes and tested for their performance to reduce CO2. Artificial biofilms were printed successfully on the copper, stainless-steel, and titanium meshes by entrapping the bacterial cells on the alginate-based hydrogel. S. ovata has been shown to reduce CO2 to organic compounds using either electrons supplied by the cathode or hydrogen produced at the cathode.

3.1. MES Performance of Bioprinted Cu Mesh

Two MES reactors with S. ovata bioprinted copper cathodes, namely CuBP1 and CuBP2, were operated as batch reactors at −0.8 V vs. Ag/AgCl. High reduction currents ranging from −11 mA to −28 mA were recorded at the starting phase of both copper bioprinted (CuBP) cathodes (especially in the first 5 days of CuBP1) (Figure 2A). This phenomenon of high current at the starting phase was also mentioned by Chatzipanagiotou et al. [35,36] while using Cu and Ni as cathodes. Electrochemical reduction of metal oxides could be associated with the initial high currents [23,35]. Annealing of Cu produced a uniform layer of CuOx (Figure 3A,B) on the electrode surface that was expected to resist the possible corrosion of Cu and to some extent prevented the anti-bacterial effect of Cu ions. In CuBP1 MES, the current started to decrease after 5–10 days and then stabilized around −5 to −8 mA (Figure 2A), whereas the stabilized current in CuBP2 was around −2 mA at −0.8 V (Supplementary Materials Figure S1).
The outer layer of bioprint layer started to fall off the cathode over time. The MES operation was continued, as a thin layer of bioprint, along with the encapsulated bacteria, remained attached. The shredded bioprint bacteria may also have contributed to CO2 reduction, being in suspended form. Immobilized cells on the surface of the alginate bioprint are visible in SEM images of the CuBP1 cathode (Figure 3C–E). In bioprinting, the hydrogel containing alginate and methacrylate was crosslinked using low-intensity UV radiation after printing on the carbon cloth. After crosslinking, the bioprint layer becomes mechanically robust and chemically resistant. Furthermore, the microbial biodegradation of gelatin is rare and has not been reported for S. ovata. It is therefore unlikely for the bioprint to contribute as a carbon source while operating as biocathode. However, the breakage of bioprint from the cathode was observed, which was likely related to the adhesion on the metal, rather than microbial degradation. Although the bioprinted biofilms were sufficiently stable for the duration of the experiments, the alginate hydrogels were not ideal for long-term operation on cathodes. MES cathodes typically experience a localized increase in pH at the biofilm–cathode interface, resulting in pH values far above the bulk pH [37]. This increased pH might result in faster alginate detachment, since alginate is known to degrade at pH values above 10 [38]. This might have led to an increase in the breaking rate of the bioprinted biofilm at the cathode–hydrogel interface, resulting in the biofilm sloughing off over time. Although the alginate hydrogels are not ideally suitable for long-term cathodic biofilms, we believe that the results would still be indicative of bioprinted MES performance, since the gels remained attached for a sufficient duration. Further experiments are suggested using more stable, chemically crosslinked hydrogels, such as UV-crosslinked or gelatin-based hydrogels.
The concentration of acetate increased steadily in both CuBP1 and CuBP2 over the first 8 days, reaching around 0.3 to 0.5 g/L, but in CuBP2 the concentration did not increase after 0.3 g/L, but rather started to decline, although high cathodic currents were recorded (Figure 2B). Due to this low production, we are not discussing CuBP2 further. The combined effect of copper leaching and mass-transfer limitation of substrates to the bacteria within the bioprint could be the main reason that triggered the inactivation of bacteria and thus resulted a drop in current in CuBP2 after eight days of operation.
The highest acetate production rate at this initial period of operation in CuBP1 was 0.07 ± 0.01 g/L/d (30.2 ± 5.4 g/m2/d) with a coulombic efficiency of 26%. A small amount of formate production (up to 640 mg/L) was also observed in CuBP reactors, but it did not show a steady accumulation, hence it is not presented or discussed here further. To allow more hydrogen evolution at the cathode, the potentials were switched to −0.9 and −1 V in both CuBP1 and CuBP2 after 10–12 days of operation at −0.8 V. The more negative cathode potential resulted in the production of approximately 1 g/L acetate in 18 days in CuBP1. Acetate production in CuBP1 continued at a slower rate, 0.04 ± 0.02 g/L/day (16 ± 10 g/m2/d), reaching 1.74 g/L after 72 days of operation, even though a large reduction current was observed. From the acetate production profile in Figure 2B, it appears that the acetate accumulated in the CuBP1 MES only corresponded to a minority of the reduction current recorded in the MES, since the acetate equivalent calculated from the current increased much faster, theoretically supporting an acetate concentration up to 9 g/L within 40 days. Overall, the bioprinted Cu cathode produced on average 0.05 ± 0.02 g/L/day (23 ± 7 g/m2/d) acetate production. However, the MES performances of Cu cathodes remained inconsistent, as there were variations in performance with the batch operation of each individual reactor and also between the two reactors. Due to the complexities of cathode reactions, current fluctuations were observed. Indeed, the bioprinting on the metal cathode added complexities at the cathode, due to the combined effects of the electrochemical activity of the metals, mass-transfer limitation within the bioprint, and the bacterial electrochemistry. The combined effects from these multiple interactions created unknown variables at the cathode biofilm. In addition, from practical experience on electrochemical systems, reference electrodes may lose their stability during long-term operation. The changes in current in the reactors can also be attributed to the fluctuations in individual reference electrodes. So, the duplicate reactors showed different currents when the status of individual reference electrodes changes.
The copper oxides deposited on the cathodes are known for their electrochemical catalysis of CO2 reduction, and the combination of CuOx and microbial catalyst was reported to enhance CO2 reduction [23,35]. Layers of copper oxide formed after annealing at 300 °C were visible on the SEM image of Figure 3B. In our study, the bioprinted Cu cathodes did not show a high rate of CO2 reduction, most likely due to microbial inhibition caused by Cu ions. Transient metal electrode leaching cannot be avoided while the electrode is submerged in a strong electrolyte without applying reduction potential. Chatzipanagiotou et al. [23,35] showed an increase in the Cu ion concentration of the catholyte at the beginning of batch operation when using Cu cathodes in MES. Over time, while maintaining the cathodic electrical potentials, the concentration of Cu ions slowly declined because the metal ions were reduced and deposited on the cathode [23,35]. Aryal et al. [33] also showed that S. ovata did not thrive on a copper-coated graphene cathode, whereas high performance was observed when graphene was coated on top of copper foam. These studies suggest that direct contact of S. ovata with the surface of less corrosion-resistant metals such as copper could be inhibitory to the bacteria under the electrochemical reactor conditions. A comparatively low performance of S. ovata observed in CuBP reactors might be due to the similar conditions in this study. Corrosion-like features were visible on the SEM image (Figure 3F) of the CuBP1 cathode. However, further analyses regarding the Cu ion concentration and the impact of solid Cu on S. ovata growth were not carried out in this study.

3.2. MES Performance of Bioprinted SS Mesh

Bioprinting of S. ovata was carried out on a stainless-steel (SS) mesh named SSBP, without the annealing as mentioned for Cu mesh. The acetate concentration and reduction currents from SSBP are presented in Figure 4A,B. Stainless steels are known for chemical stability due to passive layer of oxide formation [39], and SSs are commonly used in MES reactors as a current collector [40]. SS cathodes are known low-cost non-noble electrocatalysts for hydrogen evolution [41,42]. As soon as the bioprinted SS cathodes were polarized at −0.8 V vs. Ag/AgCl to start MES operation, a high current of −4 mA was observed (Figure 4A). The high current resulted in a high rate of H2 evolution through water electrolysis. In addition, the reduction of the oxide layer formed on the SS could be the reason for the high initial current, as discussed previously. Within a day, the current decreased to −1 mA, most likely because the available metal oxides might have been reduced. The reduction current in the SSBP MES reactors increased over several days and remained between −5 to −10 mA (Figure 4A). As described earlier for CuBP MES, a portion of the bioprint layer on the SS mesh broke off and a thin layer of bioprinted gel remained on the SS cathodes. SEM images in Figure 4C,D show that the SS cathode was covered with a considerably thick biofilm. Acetate levels in the catholyte started to rise from the beginning when the cathode was polarized at −0.8V vs. Ag/AgCl. The production continued steadily and reached 0.74 g/L of acetate in 8 days (Figure 4B). This acetate level corresponds to a highest production rate of 38.62 ± 1.4 g/m2/d and SSBP showed continuous accumulation of acetate afterward, maintaining an average rate of 0.1 ± 0.01 g/L/d (24.6 ± 8.8 g/m2/d). The sizes of SS and Cu cathodes used in this study were relatively small (2.5 × 2.5 cm) for the reactor volume (270 mL), thus the volumetric acetate production rates were relatively low in these MES reactors, as compared to the rates obtained with bioprinted carbon cloth in Krige et al. [16].
The reduction currents recorded in the SSBP reactor were relatively low up to 30 days of operation, and thus the acetate equivalents calculated from the currents were lower than the measured acetate (Figure 4A,B). Higher acetate measurements than electrical equivalents indicated the presence of additional electron sources other than electricity during that operation period; thus, the coulombic efficiency exceeded 100% for that period. Although the additional electron sources were not identified, the contribution from cell decay is most probable. A small amount (0.5 g/L) of yeast extract was also available for bacterial growth, which may partly contribute as a carbon source for acetate production. Nevertheless, the contribution remained minimal, since the acetate accumulation in the electrolyte reached 2.6 g/L (significantly higher than the 0.5 g/L yeast extract) over the course of MES operation, confirming that CO2 was fixed to produce acetate. After 30 days of operation, higher currents were observed and acetate equivalents from electricity exceeded the measured acetate level. The accumulation of slightly more than 2.5 g/L of acetate corresponds to 51% of CE over 70 days of operation of SSBP MES.

3.3. MES Performance of Bioprinted Ti Mesh

After bioprinting of S. ovata on Ti mesh, the MES, named TiBP, was operated at −0.8 V vs. Ag/AgCl cathode potential. A high current of −25 to −30 mA was recorded in TiBP MES accompanying a high rate of H2 evolution through water electrolysis and acetate production (Figure 5A,B). An additional Ti cathode was previously used in an MES to support high-rate CO2 reduction [21]. The reduction of oxides formed on the Ti also contributed to the high reduction current. Ti is a known HER catalyst, with high resistance to corrosion due to the passive oxide layer developed on its outer surface [43]. The acetate concentration in the catholyte quickly increased at the early stage of operation, reaching the highest instantaneous production rate of 0.84 g/L/d (130.5 g/m2/d). The acetate production stabilized after 4 days, and remained around 0.34 g/L/d (53 g/m2/d) on average for 18 days, reaching a concentration of 4 g/L acetate. After day 18, the acetate production continued steadily at an average rate of 0.1 g/L/d for a long time, up to 87–90 days, reaching 9.4 g/L at the end of the run (Figure 5A). The CE of acetate production in the TiBP reactor ranged from 75 to 35% over a long operation of 90 days.
What is noteworthy in TiBP is that the bioprint gel layer did not break and remained attached to the Ti mesh, despite the fact that high currents and hydrogen evolution were observed. SEM images of the Ti cathode in Figure 6A,B show a full coverage of the Ti surface with the biofilm. So far, it appeared that the bonding of alginate gel with Ti or with its oxide was stronger than bonding with other two metals. In addition, the pH gradient developed at the electrode–bioprint interface was not severe to the bioprint layer when using Ti. Ti is resistant to corrosion, catalytic to HER, and also known for high biocompatibility in comparison to SS and Cu. This property of Ti might have allowed the high stability and high performance of the bioprinted S. ovata. However, detailed analyses in this direction remained out of the scope of this study.

4. Discussion

4.1. Bioprinting of S. ovata Relatively Increased Acetate Production from CO2 Reduction

S. ovata has been used as a biocatalyst to reduce CO2 to acetate using several metal cathodes in a microbial electrosynthesis (MES) system. A comparative overview of performance of S. ovata in MES with different cathode materials is presented in Table 1. Previous studies have reported that the acetate production rate using S. ovata in MES system with graphene-functionalized carbon cloth at −0.93 V was almost 14 g/m2/day [28], and on Si nanowire array cathodes it was 44.3 g/m2/d at −1.4 V [18]. An acetate production rate of 1.85 g/m2/d was reported at −0.6 V with Ni hollow-fiber cathodes with carbon nanotube coating [20]. In a similar study of metal cathodes for methane production from CO2, Wang and coworkers employed stainless steel, copper, and nickel mesh biocathodes and reported increased methane generation with nickel, likely due to the higher electrocatalytic activity of this electrode for the synthesis of H2 [44].
The bioprinting technique of S. ovata has been shown to effectively localize a high density of bacteria on the cathode [16]. The results from previous publications, Krige et al. [16] and Bajracharya et al. [21], on 3D bioprinting were used as a control study for the present metal bioprinting experiments. In Krige et al. [16] plain carbon cloth was used, whereas in this study the carbon cloth electrodes were replaced with the metal meshes and the same procedure for bioprinting was followed. The comparison of results of this study to our previous studies are presented in Figure 6B and discussed here. While bioprinting on plain carbon cloth, Krige et. al. [16] achieved a maximum of 52 g/m2/d, whereas in this present study, the acetate production rate was increased to 72 g/m2/d (maximum) with the Ti cathodes and bioprinting. With Cu and SS cathodes, we observed less than half the acetate production rate than Ti with bioprinted S. ovata in this study, and the rates from Cu and SS were lower than the rates from bioprinting of S. ovata on carbon cloth as reported in Krige et al. [16]. An important experimental consideration to note here in this study is that the size of the Ti cathode was three times bigger than CuBP and SSBP. The total current recorded in TiBP was relatively higher than CuBP and SSBP due to the higher area of Ti. To compare the currents with different bioprinted cathodes, the results of linear sweep voltammetry (LSV) carried out on the different cathodes at 5 mV per second scan rate are shown in Figure 7. CuBP showed the lowest onset potential (~−0.5 V vs. Ag/AgCl) for cathodic reaction in Figure 7. MES with Ti reached similar current densities as Cu, but the onset potential for cathodic reaction was between −0.6 to −0.7 V vs. Ag/AgCl. Current density from Ti was higher than that of Cu after −0.9 V vs. Ag/AgCl. The HER performances of CuBP and TiBP were higher than that of SSBP and non-bioprinted C-cloth, as the current densities for SSBP and C-cloth are considerably lower in Figure 7. The current densities are normalized to the projected area of the cathode. At more negative potentials than −0.8 V vs. Ag/AgCl, the current densities from TiBP are highest, then those from CuBP and the lowest from SSBP. For visual comparison purposes, the LSV from non-bioprinted C cloth obtained from Krige et al. [16] is also shown. In previous study [21], also it was also reported that a larger area Ti cathode with S. ovata had shown lower currents than that from the graphite rod cathode when highly active S. ovata biofilm was established on it. Thus, the high currents with the Ti cathode were not only due to the area but also due to the higher activity of bacteria on the Ti cathode as evident from Figure 7. In fact, electrochemical catalysis is actually dependent on the active surface area of the electrode rather than the physical area of the electrode [45]. Acetate production rates are also presented by normalizing to the electrode area (per m2) for the comparison of performances of different metal cathodes in Table 1. The electrode area-based production rates allow for the comparison of different cathodes’ performances although they differ in sizes, and is a standard method for reporting MES productivity. The area-based acetate production rates were more than doubled in TiBP compared to the rates from CuBP and SSBP (Table 1).
The use of metal cathodes and the bioprinting of microbes combine the benefit of direct metal electrocatalysis for H2 production and the subsequent reduction of CO2 by microorganisms, using H2 as a substrate for biological reactions, improving the overall production rate. Krige et al. [16] introduced 3D bioprinting of S. ovata on carbon cloth for MES. As a continuation, in this study we used 3D bioprinting on several metal mesh cathodes. Bioprinting allows the bacteria to be localized on the cathode rather than being in suspension and slowly forming cathode biofilm at their own pace, if at all; thus, bioprinting increased the possibilities of bacterial interaction with the cathode [16]. Fast start-up and bacterial localization on the cathode for immediate access to the reducing equivalents are the main advantages of bioprinting on the cathode. These advantages improved the CO2 reduction rates in our MES studies. Additionally, the encapsulated and localized bacterial cells in the cathode surface within the hydrogel of bioprinting also appeared helpful in avoiding the inhibitory effect of minor levels of oxygen penetration from the anode side. Hence, the bioprinting technique is useful to operate MES which has an O2-evolving anode for long-term operation, as diffusion limitation is created with the bioprinting technique. However, the bioprinting technique may need further improvement and adaptation for the enhancement of biocathode for CO2 reduction in MES, as there was low reproducibility in the performances of similarly bioprinted cathodes. This study is one of the pioneering works on the use of bioprinting to form biocathodes for CO2 reduction; optimized procedures and bioink composition are still to be attained in future study.

4.2. Metal Degradation in MES and Its Effect on MES Performance

Non-noble metals on exposure to air form varying thickness of oxides layers having semiconducting properties [39,46]. These passivating oxide layers are also likely to play a major role in corrosion resistance and in electrochemical interaction between microorganisms and metals. High chloride and sulfide ion concentrations, increased temperature, and acidic pH can induce significant passive layer disintegration on metal producing metal ions [47]. Adsorption of aggressive anions has been shown to react with metal ions and produce chemical dissolution; hence, more aggressive anions trigger the disintegration of a passive layer and commence pitting. Low pH of the solution instigates constant corrosion. Metallic cathodes in MES are likely to be corroded under biotic conditions [33]. In our study, SEM images of some spots on a CuBP cathode show morphologies similar to pitting corrosion on the surface (Figure 3F). Stainless-steel mesh in SSBP did not show such morphologies, but SS mesh contains Ni which can be corroded in the electrolyte environment. Baudler et al. [48] have also indicated the incompatible interactions of microorganism with nickel-containing electrodes, showing the lower current density obtained with stainless steel (containing 8% to 10.5% w/w nickel) in comparison to graphite bioanodes.
Bacterial production of catalase enhances the microbial corrosion of metal surfaces due to an increased oxygen reduction current, oxidizing ferric into ferric oxide, resulting in a red color in the growth system [49]. Despite the fact that the cathodic redox potential should have prevented the leaching of copper during normal operation, studies have suggested that copper might be harmful to biofilms [50]. The risk of air/oxygen infiltration in the cathode chamber and probable risk of forming corrosive species can also not be eliminated. An example from Tashiro et al. [51] showed that the electrochemical conditions for water electrolysis induced the inhibitory effects of reduction products such as H2O2 formed with slight exposure to oxygen which can react with the electrode and also prolonged the lag phase for bacterial growth. However, as discussed in the previous subsection, bioprinting can minimize such an inhibitory effect arising due to the oxygen penetration from anode side to some extent. Moreover, many factors may also impair the microbe–material interaction in MES. Further study can be suggested to characterize the factors that may negatively affect the performance of bioprinted metal cathodes.

5. Conclusions

This research demonstrates the potential of using 3D bioprinting on metals to create synthetic biofilms of S. ovata for microbial electrosynthesis. When using less corrosion-resistant and less biocompatible metals such as Cu, stainless steel, the bioprinting did not enhance the microbial electrosynthesis, presumably due to the instability of the bioprint and the inhibitory effects of metal ions on the bacteria. When using bioprinted S. ovata on a corrosion-resistant and biocompatible Ti mesh cathode, a high acetate production rate of 53 ± 19 g/m2/d at −0.8 V vs. Ag/AgCl cathode potential was obtained. Bioprinting artificial biofilms on chemically stable metals can help to develop more efficient biocathodes with locally immobilized biofilm for large-scale microbial electrosynthesis systems. A synthetic biofilm can also improve stability, allowing continuous operation without the risk of cell washing-out.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/fermentation10010034/s1, Figure S1: Profiles of cathode potentials and electric current in CuBP2 MES.

Author Contributions

Conceptualization, S.B. and A.K.; methodology, S.B.; software, S.B.; validation, S.B., A.K. and L.M.; formal analysis, S.B.; investigation, S.B. and A.K.; resources, S.B. and A.K.; data curation, S.B.; writing—original draft preparation, S.B. and A.K.; writing—review and editing, S.B., A.K. and L.M.; visualization, S.B.; supervision and funding acquisition, P.C. and U.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Swedish Research Council (VR), project numbers 2017-04867 and 2018-03875.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data are contained within the article and supplementary materials.

Acknowledgments

Authors would like to thank Kateřina Hrůzová (Luleå University of Technology) for carrying out the scanning electron microscopy (SEM) imaging.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Nevin, K.P.; Woodard, T.L.; Franks, A.E.; Summers, Z.M.; Lovley, D.R. Microbial Electrosynthesis: Feeding Microbes Electricity to Convert Carbon Dioxide and Water to Multicarbon Extracellular Organic. MBio 2010, 1, e00103-10. [Google Scholar] [CrossRef] [PubMed]
  2. Nevin, K.P.; Hensley, S.A.; Franks, A.E.; Summers, Z.M.; Ou, J.; Woodard, T.L.; Snoeyenbos-West, O.L.; Lovley, D.R. Electrosynthesis of organic compounds from carbon dioxide is catalyzed by a diversity of acetogenic microorganisms. Appl. Environ. Microbiol. 2011, 77, 2882–2886. [Google Scholar] [CrossRef] [PubMed]
  3. Prévoteau, A.; Carvajal-Arroyo, J.M.; Ganigué, R.; Rabaey, K. Microbial electrosynthesis from CO2: Forever a promise? Curr. Opin. Biotechnol. 2020, 62, 48–57. [Google Scholar] [CrossRef] [PubMed]
  4. Jourdin, L.; Sousa, J.; van Stralen, N.; Strik, D.P.B.T.B. Techno-economic assessment of microbial electrosynthesis from CO2 and/or organics: An interdisciplinary roadmap towards future research and application. Appl. Energy 2020, 279, 115775. [Google Scholar] [CrossRef]
  5. Wood, J.C.; Grové, J.; Marcellin, E.; Heffernan, J.K.; Hu, S.; Yuan, Z.; Virdis, B. Strategies to improve viability of a circular carbon bioeconomy—A techno-economic review of microbial electrosynthesis and gas fermentation. Water Res. 2021, 201, 117306. [Google Scholar] [CrossRef] [PubMed]
  6. Tremblay, P.-L.; Höglund, D.; Koza, A.; Bonde, I.; Zhang, T. Adaptation of the autotrophic acetogen Sporomusa ovata to methanol accelerates the conversion of CO2 to organic products. Sci. Rep. 2015, 5, 16168. [Google Scholar] [CrossRef] [PubMed]
  7. Zhang, T.; Nie, H.; Bain, T.S.; Lu, H.; Cui, M.; Snoeyenbos-West, O.L.; Franks, A.E.; Nevin, K.P.; Russell, T.P.; Lovley, D.R. Improved cathode materials for microbial electrosynthesis. Energy Environ. Sci. 2013, 6, 217. [Google Scholar] [CrossRef]
  8. Aryal, N.; Halder, A.; Zhang, M.; Whelan, P.R.; Tremblay, P.L.; Chi, Q.; Zhang, T. Freestanding and flexible graphene papers as bioelectrochemical cathode for selective and efficient CO2 conversion. Sci. Rep. 2017, 7, 9107. [Google Scholar] [CrossRef]
  9. Aryal, N.; Tremblay, P.-L.; Xu, M.; Daugaard, A.E.; Zhang, T. Highly Conductive Poly(3,4-ethylenedioxythiophene) Polystyrene Sulfonate Polymer Coated Cathode for the Microbial Electrosynthesis of Acetate from Carbon Dioxide. Front. Energy Res. 2018, 6, 72. [Google Scholar] [CrossRef]
  10. Marshall, C.W.; Ross, D.E.; Fichot, E.B.; Norman, R.S.; May, H.D. Electrosynthesis of commodity chemicals by an autotrophic microbial community. Appl. Environ. Microbiol. 2012, 78, 8412–8420. [Google Scholar] [CrossRef]
  11. Jourdin, L.; Freguia, S.; Donose, B.C.; Chen, J.; Wallace, G.G.; Keller, J.; Flexer, V. A novel carbon nanotube modified scaffold as an efficient biocathode material for improved microbial electrosynthesis. J. Mater. Chem. A 2014, 2, 13093–13102. [Google Scholar] [CrossRef]
  12. Jourdin, L.; Winkelhorst, M.; Rawls, B.; Buisman, C.J.N.; Strik, D.P.B.T.B. Enhanced selectivity to butyrate and caproate above acetate in continuous bioelectrochemical chain elongation from CO2: Steering with CO2 loading rate and hydraulic retention time. Bioresour. Technol. Rep. 2019, 7, 100284. [Google Scholar] [CrossRef]
  13. Jourdin, L.; Raes, S.M.T.; Buisman, C.J.N.; Strik, D.P.B.T.B. Critical Biofilm Growth throughout Unmodified Carbon Felts Allows Continuous Bioelectrochemical Chain Elongation from CO2 up to Caproate at High Current Density. Front. Energy Res. 2018, 6, 7. [Google Scholar] [CrossRef]
  14. LaBelle, E.V.; May, H.D. Energy Efficiency and Productivity Enhancement of Microbial Electrosynthesis of Acetate. Front. Microbiol. 2017, 8, 756. [Google Scholar] [CrossRef] [PubMed]
  15. Bajracharya, S.; Vanbroekhoven, K.; Buisman, C.J.N.; Strik, D.P.B.T.B.; Pant, D. Bioelectrochemical conversion of CO2 to chemicals: CO2 as a next generation feedstock for electricity-driven bioproduction in batch and continuous modes. Faraday Discuss. 2017, 202, 433–449. [Google Scholar] [CrossRef] [PubMed]
  16. Krige, A.; Rova, U.; Christakopoulos, P. 3D bioprinting on cathodes in microbial electrosynthesis for increased acetate production rate using Sporomusa ovata. J. Environ. Chem. Eng. 2021, 9, 106189. [Google Scholar] [CrossRef]
  17. Kong, F.; Ren, H.Y.; Liu, D.; Wang, Z.; Nan, J.; Ren, N.Q.; Fu, Q. Improved decolorization and mineralization of azo dye in an integrated system of anaerobic bioelectrochemical modules and aerobic moving bed biofilm reactor. Bioresour. Technol. 2022, 353, 127147. [Google Scholar] [CrossRef] [PubMed]
  18. Su, Y.; Cestellos-Blanco, S.; Kim, J.M.; Shen, Y.; Kong, Q.; Lu, D.; Liu, C.; Zhang, H.; Cao, Y.; Yang, P. Close-Packed Nanowire-Bacteria Hybrids for Efficient Solar-Driven CO2 Fixation. Joule 2020, 4, 800–811. [Google Scholar] [CrossRef]
  19. Nie, H.; Zhang, T.; Cui, M.; Lu, H.; Lovley, D.R.; Russell, T.P. Improved cathode for high efficient microbial-catalyzed reduction in microbial electrosynthesis cells. Phys. Chem. Chem. Phys. 2013, 15, 14290–14294. [Google Scholar] [CrossRef]
  20. Bian, B.; Alqahtani, M.F.; Katuri, K.P.; Liu, D.F.; Bajracharya, S.; Lai, Z.P.; Rabaey, K.; Saikaly, P.E. Porous Nickel Hollow Fiber Cathodes Coated with CNTs for Efficient Microbial Electrosynthesis of Acetate from CO2 using Sporomusa ovata. J. Mater. Chem. A 2018, 6, 17201–17211. [Google Scholar] [CrossRef]
  21. Bajracharya, S.; Krige, A.; Matsakas, L.; Rova, U.; Christakopoulos, P. Dual cathode configuration and headspace gas recirculation for enhancing microbial electrosynthesis using Sporomusa ovata. Chemosphere 2022, 287, 132188. [Google Scholar] [CrossRef] [PubMed]
  22. Guo, K.; Prévoteau, A.; Patil, S.A.; Rabaey, K. Engineering electrodes for microbial electrocatalysis. Curr. Opin. Biotechnol. 2015, 33, 149–156. [Google Scholar] [CrossRef] [PubMed]
  23. Chatzipanagiotou, K.; Soekhoe, V.; Jourdin, L.; Buisman, C.J.N.; Bitter, J.H.; Strik, D.P.B.T.B. Catalytic Cooperation between a Copper Oxide Electrocatalyst and a Microbial Community for Microbial Electrosynthesis. Chempluschem 2021, 86, 763–777. [Google Scholar] [CrossRef] [PubMed]
  24. Madjarov, J.; Soares, R.; Paquete, C.M.; Louro, R.O. Sporomusa ovata as Catalyst for Bioelectrochemical Carbon Dioxide Reduction: A Review across Disciplines from Microbiology to Process Engineering. Front. Microbiol. 2022, 13, 2159. [Google Scholar] [CrossRef] [PubMed]
  25. Rosenbaum, M.A.; Berger, C.; Schmitz, S.; Uhlig, R. Microbial electrosynthesis I: Pure and defined mixed culture engineering. In Advances in Biochemical Engineering/Biotechnology; Springer: Cham, Switzerland, 2019; Volume 167, pp. 181–202. Available online: https://link.springer.com/chapter/10.1007/10_2017_17 (accessed on 1 October 2023).
  26. Chen, L.; Tremblay, P.-L.; Mohanty, S.; Xu, K.; Zhang, T. Electrosynthesis of acetate from CO2 by a highly structured biofilm assembled with reduced graphene oxide–tetraethylene pentamine. J. Mater. Chem. A 2016, 4, 8395–8401. [Google Scholar] [CrossRef]
  27. Lee, C.R.; Kim, C.; Song, Y.E.; Im, H.; Oh, Y.K.; Park, S.; Kim, J.R. Co-culture-based biological carbon monoxide conversion by Citrobacter amalonaticus Y19 and Sporomusa ovata via a reducing-equivalent transfer mediator. Bioresour. Technol. 2018, 259, 128–135. [Google Scholar] [CrossRef] [PubMed]
  28. Aryal, N.; Halder, A.; Tremblay, P.L.; Chi, Q.; Zhang, T. Enhanced microbial electrosynthesis with three-dimensional graphene functionalized cathodes fabricated via solvothermal synthesis. Electrochim. Acta 2016, 217, 117–122. [Google Scholar] [CrossRef]
  29. Knoll, M.T.; Fuderer, E.; Gescher, J. Sprayable biofilm—Agarose hydrogels as 3D matrix for enhanced productivity in bioelectrochemical systems. Biofilm 2022, 4, 100077. [Google Scholar] [CrossRef]
  30. Rabaey, K.; Rozendal, R.A. Microbial electrosynthesis—Revisiting the electrical route for microbial production.pdf. Nat. Rev. Microbiol. 2010, 8, 706–716. [Google Scholar] [CrossRef]
  31. Tremblay, P.-L.; Faraghiparapari, N.; Zhang, T. Accelerated H2 Evolution during Microbial Electrosynthesis with Sporomusa ovata. Catalysts 2019, 9, 166. [Google Scholar] [CrossRef]
  32. Zeradjanin, A.R.; Grote, J.-P.; Polymeros, G.; Mayrhofer, K.J.J. A Critical Review on Hydrogen Evolution Electrocatalysis: Re-exploring the Volcano-relationship. Electroanalysis 2016, 28, 2256–2269. [Google Scholar] [CrossRef]
  33. Aryal, N.; Wan, L.; Overgaard, M.H.; Stoot, A.C.; Chen, Y.; Tremblay, P.L.; Zhang, T. Increased carbon dioxide reduction to acetate in a microbial electrosynthesis reactor with a reduced graphene oxide-coated copper foam composite cathode. Bioelectrochemistry 2019, 128, 83–93. [Google Scholar] [CrossRef]
  34. Grass, G.; Rensing, C.; Solioz, M. Metallic Copper as an Antimicrobial Surface. Appl. Environ. Microbiol. 2011, 77, 1541–1547. [Google Scholar] [CrossRef] [PubMed]
  35. Chatzipanagiotou, K.; Jourdin, L.; Buisman, C.J.N.; Strik, D.P.B.T.B.; Bitter, J.H. CO2 Conversion by Combining a Copper Electrocatalyst and Wild-type Microorganisms. ChemCatChem 2020, 12, 3900–3912. [Google Scholar] [CrossRef]
  36. Chatzipanagiotou, K.R.; Jourdin, L.; Bitter, J.H.; Strik, D.P.B.T.B. Concentration-dependent effects of nickel doping on activated carbon biocathodes. Catal. Sci. Technol. 2022, 12, 2500–2518. [Google Scholar] [CrossRef]
  37. Desloover, J.; Arends, J.B.A.; Hennebel, T.; Rabaey, K. Operational and technical considerations for microbial electrosynthesis. Biochem. Soc. Trans. 2012, 40, 1233–1238. [Google Scholar] [CrossRef] [PubMed]
  38. Pawar, S.N.; Edgar, K.J. Alginate derivatization: A review of chemistry, properties and applications. Biomaterials 2012, 33, 3279–3305. [Google Scholar] [CrossRef] [PubMed]
  39. Olsson, C.O.A.; Landolt, D. Passive films on stainless steels—Chemistry, structure and growth. Electrochim. Acta 2003, 48, 1093–1104. [Google Scholar] [CrossRef]
  40. Shakeel, S.; Khan, M.Z. Enhanced production and utilization of biosynthesized acetate using a packed-fluidized bed cathode based MES system. J. Environ. Chem. Eng. 2022, 10, 108067. [Google Scholar] [CrossRef]
  41. Liu, Y.; Zhang, K.; Zhang, D.; Dong, W.; Jiang, T.; Zhou, H.; Li, L.; Mao, B. Industrial stainless steel meshes for efficient electrocatalytic hydrogen evolution. J. Energy Storage 2021, 41, 102844. [Google Scholar] [CrossRef]
  42. Munoz, L.D.; Erable, B.; Etcheverry, L.; Riess, J.; Basséguy, R.; Bergel, A.; De Silva Munoz, L.; Erable, B.; Etcheverry, L.; Riess, J.; et al. Combining phosphate species and stainless steel cathode to enhance hydrogen evolution in microbial electrolysis cell (MEC). Electrochem. Commun. 2010, 12, 183–186. [Google Scholar] [CrossRef]
  43. Khalil, M.W.; Abdel Rahim, M.A. Hydrogen evolution reaction on Titanium and oxide-covered titanium electrodes. Materwiss. Werksttech. 1991, 22, 390–395. [Google Scholar] [CrossRef]
  44. Wang, L.; He, Z.; Guo, Z.; Sangeetha, T.; Yang, C.; Gao, L.; Wang, A.; Liu, W. Microbial community development on different cathode metals in a bioelectrolysis enhanced methane production system. J. Power Sources 2019, 444, 227306. [Google Scholar] [CrossRef]
  45. Sharma, M.; Bajracharya, S.; Gildemyn, S.; Patil, S.A.; Alvarez-Gallego, Y.; Pant, D.; Rabaey, K.; Dominguez-Benetton, X. A critical revisit of the key parameters used to describe microbial electrochemical systems. Electrochim. Acta 2014, 140, 191–208. [Google Scholar] [CrossRef]
  46. Ohmi, T.; Nakagawa, Y.; Nakamura, M.; Ohki, A.; Koyama, T. Formation of chromium oxide on 316L austenitic stainless steel. J. Vac. Sci. Technol. A Vac. Surf. Film. 1998, 14, 2505. [Google Scholar] [CrossRef]
  47. Tasić, Ž.Z.; Petrović Mihajlović, M.B.; Radovanović, M.B.; Antonijević, M.M. New trends in corrosion protection of copper. Chem. Pap. 2019, 73, 2103–2132. [Google Scholar] [CrossRef]
  48. Baudler, A.A.; Schmidt, I.; Langner, M.; Greiner, A.; Schröder, U.; Schroder, U. Does it have to be carbon? Metal anodes in microbial fuel cells and related bioelectrochemical systems. Energy Environ. Sci. 2015, 8, 2048–2055. [Google Scholar] [CrossRef]
  49. Busalmen, J.P.; Vázquez, M.; De Sánchez, S.R. New evidences on the catalase mechanism of microbial corrosion. Electrochim. Acta 2002, 47, 1857–1865. [Google Scholar] [CrossRef]
  50. Jiang, Y.; Chu, N.; Zhang, W.; Ma, J.; Zhang, F.; Liang, P.; Zeng, R.J. Zinc: A promising material for electrocatalyst-assisted microbial electrosynthesis of carboxylic acids from carbon dioxide. Water Res. 2019, 159, 87–94. [Google Scholar] [CrossRef]
  51. Tashiro, Y.; Hirano, S.; Matson, M.M.; Atsumi, S.; Kondo, A. Electrical-biological hybrid system for CO2 reduction. Metab. Eng. 2018, 47, 211–218. [Google Scholar] [CrossRef]
Figure 1. Bioprint layer of S. ovata on Cu cathode (encircled part) placed inside an MES reactor.
Figure 1. Bioprint layer of S. ovata on Cu cathode (encircled part) placed inside an MES reactor.
Fermentation 10 00034 g001
Figure 2. (A) Profiles of electric current and cathode potentials in CuBP1 MES. (B) Profiles of acetate concentration measured in CuBP1 and CuBP2 MES, and the corresponding electricity-based acetate-equivalents calculated from the recorded electric current in CuBP1.
Figure 2. (A) Profiles of electric current and cathode potentials in CuBP1 MES. (B) Profiles of acetate concentration measured in CuBP1 and CuBP2 MES, and the corresponding electricity-based acetate-equivalents calculated from the recorded electric current in CuBP1.
Fermentation 10 00034 g002
Figure 3. (A) Photograph of annealed copper meshes. (B) SEM showing the formation of copper oxides (CuOx) layers on the annealed Cu mesh. (C) Photograph of a dry bioprinted copper cathode CuBP1 after the MES operation. (D,E) SEMs showing biofilm on CuBP1. (F) Pit formation-like observation on the CuBP1 cathode after operation of MES.
Figure 3. (A) Photograph of annealed copper meshes. (B) SEM showing the formation of copper oxides (CuOx) layers on the annealed Cu mesh. (C) Photograph of a dry bioprinted copper cathode CuBP1 after the MES operation. (D,E) SEMs showing biofilm on CuBP1. (F) Pit formation-like observation on the CuBP1 cathode after operation of MES.
Fermentation 10 00034 g003aFermentation 10 00034 g003b
Figure 4. (A) Profiles of electric current and cathode potentials in bioprinted stainless-steel cathode MES (SSBP). (B) Profiles of measured acetate concentration and calculated acetate equivalents from the electric current in SSBP during the long-term operation. (C,D) SEMs showing the biofilm on the SSBP cathode at different magnifications.
Figure 4. (A) Profiles of electric current and cathode potentials in bioprinted stainless-steel cathode MES (SSBP). (B) Profiles of measured acetate concentration and calculated acetate equivalents from the electric current in SSBP during the long-term operation. (C,D) SEMs showing the biofilm on the SSBP cathode at different magnifications.
Fermentation 10 00034 g004
Figure 5. (A) Profiles of electric current and cathode potentials in TiBP MES. (B) Long-term profiles of measured acetate concentration and calculated acetate equivalents from the electric current in TiBP MES. For a comparison of performance, acetate production levels of S. ovata with carbon cloth cathodes in non-bioprinted and bioprinted MES reactors from Krige et al. [16] are also plotted, and an enlarged form of the first 10 days’ acetate production is shown on the left side.
Figure 5. (A) Profiles of electric current and cathode potentials in TiBP MES. (B) Long-term profiles of measured acetate concentration and calculated acetate equivalents from the electric current in TiBP MES. For a comparison of performance, acetate production levels of S. ovata with carbon cloth cathodes in non-bioprinted and bioprinted MES reactors from Krige et al. [16] are also plotted, and an enlarged form of the first 10 days’ acetate production is shown on the left side.
Fermentation 10 00034 g005
Figure 6. (A,B) SEMs showing the biofilm on the TiBP at different magnifications.
Figure 6. (A,B) SEMs showing the biofilm on the TiBP at different magnifications.
Fermentation 10 00034 g006
Figure 7. Linear sweep voltammetry (LSV) carried out at 5 mV per second scan rate on CuBP, SSBP, and TiBP with bioprinted S. ovata, showing a comparison of current densities from different cathodes with normalization to projected area. LSV plot from non-bioprinted carbon cloth cathode is provided for a visual impression of higher current densities in metal cathodes.
Figure 7. Linear sweep voltammetry (LSV) carried out at 5 mV per second scan rate on CuBP, SSBP, and TiBP with bioprinted S. ovata, showing a comparison of current densities from different cathodes with normalization to projected area. LSV plot from non-bioprinted carbon cloth cathode is provided for a visual impression of higher current densities in metal cathodes.
Fermentation 10 00034 g007
Table 1. A comparative overview of performances of S. ovata in MES with different cathode materials.
Table 1. A comparative overview of performances of S. ovata in MES with different cathode materials.
StudyCathodeCathode PotentialVolume-Based ProductivityArea-Based ProductivityCE for AcetateMax Product Titer
[V vs. Ag/AgCl][g/L/d][g/m2/d][%][mM]
This studySynthetic biofilm on Ti−0.80.34 ± 0.1253 ± 1955 ± 20.1157
This studySynthetic biofilm on SS−0.8 to −10.1 ± 0.0124.6 ± 8.85143
This studySynthetic biofilm on Cu−0.8 to −0.90.05 ± 0.0123.05 ± 7.12629.5
Krige et al. [16]Synthetic biofilm CC−0.80.31 ± 0.0347.3 ± 5.162.7 ± 15.460
tSu et al. [18]Si nanowire array−1.40.3044.3~80-
Bian et al. [20]Ni-PHF/CNT + direct CO2 supply−0.60.021.85832.7
Aryal et al. [28]3D graphene-functionalized CF−0.930.1213.9 ± 0.486.524
Abbreviations: SS = Stainless steel; CC = carbon cloth; CF = carbon felt; PHF = porous nickel hollow fiber; CNT = carbon nanotubes.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bajracharya, S.; Krige, A.; Matsakas, L.; Rova, U.; Christakopoulos, P. Microbial Electrosynthesis Using 3D Bioprinting of Sporomusa ovata on Copper, Stainless-Steel, and Titanium Cathodes for CO2 Reduction. Fermentation 2024, 10, 34. https://doi.org/10.3390/fermentation10010034

AMA Style

Bajracharya S, Krige A, Matsakas L, Rova U, Christakopoulos P. Microbial Electrosynthesis Using 3D Bioprinting of Sporomusa ovata on Copper, Stainless-Steel, and Titanium Cathodes for CO2 Reduction. Fermentation. 2024; 10(1):34. https://doi.org/10.3390/fermentation10010034

Chicago/Turabian Style

Bajracharya, Suman, Adolf Krige, Leonidas Matsakas, Ulrika Rova, and Paul Christakopoulos. 2024. "Microbial Electrosynthesis Using 3D Bioprinting of Sporomusa ovata on Copper, Stainless-Steel, and Titanium Cathodes for CO2 Reduction" Fermentation 10, no. 1: 34. https://doi.org/10.3390/fermentation10010034

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop