Abstract
One-carbon (C1) compounds found in greenhouse gases and industrial waste streams are underutilized carbon and energy sources. While various biological and chemical means exist for converting C1 substrates into multi-carbon products, major challenges of C1 conversion lie in creating net value. Here, we review metabolic strategies to utilize carbon across oxidation states. Complications arise in biochemical C1-utilization approaches because of the need for cellular energy currency adenosine triphosphate (ATP). ATP supports cell maintenance and proliferation and drives thermodynamically challenging reactions by coupling them with ATP hydrolysis. Powering metabolism through substrate cofeeding and energy transduction from light and electricity improves ATP availability, relieves metabolic bottlenecks, and upcycles carbon. We present a bioenergetic, engineering, and technoeconomic outlook for bringing elements to life.
Keywords: Carbon Upcycling, One-Carbon Compounds, Metabolic Engineering, Systems Biology
Motivation and Challenges for Upcycling One-Carbon Compounds
One-carbon (C1) compounds are gaining traction as sustainable feedstocks to produce valuable chemicals. C1 compounds are often found in waste streams, and their conversion to value-added products offers environmentally beneficial alternatives to conventional chemical processes, which start from fossil resources and emit greenhouse gases. Methane and carbon dioxide (CO2) are greenhouse gases which pose a global threat. Carbon monoxide (CO) is a poisonous gas found in syngas and partial combustion. Formate, formaldehyde, and methanol are C1 intermediates of methane, CO, and CO2 conversion. Various technologies for utilizing C1 substrates exist today. Electrolysis uses electric potential to push redox reactions that are otherwise non-spontaneous. By applying electric potential, CO2 can be converted to other C1 products [1, 2] and, with serial reduction, simple multi-carbon products such as acetate, ethanol and ethylene [3, 4]. Bioconversion, on the other hand, uses natural and synthetic pathways consisting of enzyme-catalyzed reaction steps to convert C1 feedstocks into biomass and a plethora of bioproducts.
Cellular metabolism encompasses biochemical reactions responsible for the synthesis of biomass and bioproducts as well as the generation of cellular energy currency adenosine triphosphate (ATP) and reducing power. Metabolic engineering endeavors rewire cellular metabolism to confer organisms with new and improved bioconversion capabilities. Many C1-utilizing organisms exist in nature, and their metabolism has been mapped [5–8]. Synthetic and systems biology tools have contributed to understanding and harnessing the full potential of metabolic pathways [9]. Different metabolic pathways are distinctively suitable for accommodating substrates and products with bioenergetic feasibility and biochemical proximity in terms of the number of reactions steps. Thus, for efficient and economically viable bioconversion, metabolic engineering is accompanied by a careful pairing of substrate and product [10].
Like any other system, operating metabolism costs energy. Thus, we asked three questions: Where does energy come from? How can energy be supplied without interruption? Does the output value exceed the input energy cost? Fortunately, cells can power metabolism using chemical, electrical, and light energy. Nutrient environments dictate how cells generate usable energy ATP and reducing equivalents from organic molecules. Metabolism works in synergy with abiological systems such as electrochemical cells and light-harvesting materials. Nonetheless, creating value while upcycling C1 compounds remains challenging. Here, we outline the pathways of one-carbon metabolism, diverse energy provision strategies, and biochemical and energetic considerations for efficient C1 conversion. Energy diversification and efficiency maximization principles will engender market-favored green technologies that outcompete conventional chemical processes.
Main Text
Physicochemical Considerations for C1 Compound Utilization
Efficient conversion of C1 compounds goes hand in hand with efficient energy transduction. Comparison of the oxidation states between C1 substrates and desired products indicates how many electrons need to be added or moved for the conversion. In electrolysis, an electric potential is applied to provide a driving force for electron transfer. In biology, enzymes interconvert and assimilate C1 substrates (Figure 1a). While oxidation number does not capture the thermodynamic, kinetic, and mass transport phenomena [2, 3, 11], it alludes to potential intermediates and side products. It also indicates whether a substrate can be an energy source for living organisms. For example, carbon in its most oxidized state, CO2, can be building blocks for biomolecules, but energy needs to be supplied to cells for the assimilation of CO2 into metabolism.
Figure 1. Energetics of different one-carbon substrates.
(a) Comparison between the oxidation numbers of one-carbon (C1) substrates and the average oxidation numbers of the carbons of metabolites and bioproducts. The fatty acid oxidation numbers range from 0 (acetic acid, a short-chain fatty acid) to ~2 (very-long-chain fatty acids). The amino acids oxidation number indicates the average of the carbons of the 20 canonical amino acids [94]. Enzymatic reactions that facilitate interconversion between C1 substrates include: MMO, methane monooxygenase; MDH, methanol dehydrogenase; MOX, methanol oxidase; FDH, formate dehydrogenase. (b) Comparison of the oxidation number of the carbon and the number of known major pathways for each substrate. (c) Comparison of the oxidation number of the carbon and number of ATP molecules required to produce convergent central carbon metabolites on a per-carbon basis for each C1 substrate. CH4 represents methane; CH3OH, methanol; CH2O, formaldehyde; CO2, carbon dioxide; HCOOH, formic acid; MMO, methane monooxygenase; MDH, methanol dehydrogenase; MOX, methanol oxidase; CODH, carbon monoxide dehydrogenase; FADH, formaldehyde dehydrogenase; FDH, formate dehydrogenase; RuMP, ribulose monophosphate cycle; EuMP, erythrulose monophosphate cycle; FORCE, formyl-CoA elongation reactions; HSC, homoserine cycle; WLP, Wood-Ljungdahl pathway; rGlyP, reductive glycine pathway; SerC, Serine Cycle; and CBB, Calvin-Benson-Bassham cycle.
C1 substrates can be assimilated via several metabolic pathways (Figure 1b) with varying ATP requirements (Figure 1c). The presence of C1 assimilation pathways does not immediately render them feasible alternatives to conventional chemical processes. Despite the Wood-Ljungdahl pathway (WLP) in acetogenic bacteria for converting CO2, CO, and formate into acetate and downstream commodity chemicals, substantial bioenergetic constraint exists [12]. During gas fermentation, acetogens are ATP-limited due to diffusion and solubility limitations as well as a lack of substrate-level phosphorylation [13]. Thus, ATP generation is a key engineering objective for biochemical C1 upcycling.
Biochemical Considerations for Pairing C1 Compounds and Sensible Bioproducts
From a biological perspective, an ideal C1 substrate provides the carbon backbone for not only bioproducts but also biomass precursors (e.g., amino acids, nucleotides, and lipids). One-carbon metabolism condenses C1 substrates to two-carbon (C2) and/or three-carbon (C3) metabolites. Each C1 metabolic pathway has a unique cofactor (e.g., ATP and redox pairs) requirement and thermodynamics (Table 1) for assimilation into central carbon metabolism (Figure 2). From the resulting C2 and C3 metabolites, sensible bioproducts to synthesize are those that avoid as much oxidation and decarboxylation as possible.
Table 1:
C1 Metabolic pathway summary including substrates required and resulting products, examples of organisms in which the pathway occurs naturally, and Gibbs free energy of the overall reaction (ΔG’°).
Pathway | Substrates | Products | Relevant organisms | ∆G’° (kJ/mol) |
References |
---|---|---|---|---|---|
Wood-Ljungdahl Pathway (WLP) | 4H2 + 2CO2 | CH3COOH + 2H2O | Acetogens e.g., Moorella thermoacetica, Clostridium ljungdahlii, Acetobutylicum woodii, etc. |
−95 | [5] |
Calvin Cycle (CBB) to Glucose | 6CO2 + 12NADPH + 12H2O + 18ATP | C6H12O6 + 12NADP+ + 18ADP + 18Pi | Plants, cyanobacteria, algae, etc. | −358.8 | [91] |
Calvin Cycle (CBB) | 3CO2 + 6NADPH + 9ATP + 5H2O | GAP + 6NADP + 9ADP + 8Pi | Plants, cyanobacteria, algae, etc. | −141.8 | [92] |
Reductive Glycine Pathway (rGlyP) | CHOOH + ATP + (THF) + 2NADH + NH3 + CO2 | ADP + Pi + H2O + 2NAD+ + (THF) + C2H5NO2 | Desulfovibrio desulfuricans, Saccharomyces cerevisiae, etc. | −17.2 | [7, 29] eQuilibrator (equilibrator. weizmann.ac.il) |
Serine Cycle | HCHO + CO2 + 2ATP + 2NADH + CoA | Acetyl-CoA + 2ADP + 2Pi + 2NAD+ | Type II Methanotrophs e.g., Hyphomicrobium methylovorum GM2, Methylobacter whittenburyk, Methylocystis parvus, etc. |
−123.5 | [8, 40, 93] eQuilibrator (equilibrator. weizmann.ac.il) |
Ribulose Monophosphate Cycle (RuMP) | 3CH2O + ATP | DHAP + ADP | Type I Methanotrophs e.g., Bacilus subtilis, Methylobacillus flagellates, Mycobacterium gastri MN19, etc. |
−89.9 | MetaCyc (MetaCyc.org), eQuilibrator (equilibrator. weizmann.ac.il) |
Figure 2. Assimilation of C1 compounds into central metabolism.
The oxidation states of C1 compounds and products are shown along metabolic reaction steps and pathways. CH4 represents methane; CH3OH, methanol; CH2O, formaldehyde; CO2, carbon dioxide; HCOOH, formic acid; CH3COOH, acetic acid; C2H5NO2, glycine; C2H4O3, glycolic acid; MMO, methane monooxygenase; MDH, methanol dehydrogenase; MOX, methanol oxidase; CODH, carbon monoxide dehydrogenase; FADH, formaldehyde dehydrogenase; FDH, formate dehydrogenase; RuMP, ribulose monophosphate cycle; EuMP, erythrulose monophosphate cycle; FORCE, formyl-CoA elongation reactions; WLP, Wood-Ljungdahl pathway; rGlyP, reductive glycine pathway; and CBB, Calvin-Benson-Bassham cycle.
Photosynthetic organisms including plants and algae, the most familiar assimilators of CO2, use the Calvin-Benson-Bassham (CBB) cycle to generate multi-carbon intermediates (Figure 3a) [14, 15]. The CBB cycle also enables utilization of CO in organisms that express carbon monoxide dehydrogenases, which may simultaneously provide electrons for respiration [16]. Interestingly, the genes for the CBB cycle are found in 6% of microbes whose genomes are sequenced [6], suggesting it could be engineered in non-photosynthetic organisms. Despite its relative prominence, the CBB cycle requires 3 ATP and 2 NADPH for reducing each CO2 molecule. Ribulose bisphosphate carboxylase/oxygenase (RuBisCO) catalyzes the carboxylation step in the CBB cycle, but due to its slow turnover, it remains a bottleneck for the wider application of the pathway [6]. Introducing CBB genes to tractable model organisms may increase carbon yield by limiting net CO2 loss. Heterologous expression of RuBisCO and phophoribulokinase (PRK) in E. coli in conjunction with its central carbon metabolism forms the CBB cycle, but regulatory and kinetic effects must also be considered for the CBB cycle to turn and fix CO2 [17]. The expression of CBB proteins in E. coli can be negatively impacted by protein aggregates termed inclusion bodies (IB), which result from overexpressing foreign or mutated genes. IB formation is sensitive to environmental conditions; it has been shown that reducing the culture temperature from 37°C to 30°C results in uninterrupted CBB-cycle protein expression and increases CO2 recycling to glyceraldehyde-3-phosphate (GAP) and subsequently lower glycolytic intermediates. Reducing the culture temperature of the CBB-capable E. coli strain led to a 2.3-fold increase in pyruvate production [18]. Using the C3 product of the CBB cycle to synthesize bioproducts that are derived from downstream C2 metabolites (e.g., acetyl-CoA and acetaldehyde) rather than C3 metabolites would lower carbon yield.
Figure 3. Major C1 assimilation pathways and potential products.
(a) The Calvin-Benson-Bassham Cycle (CBB) converts three CO2 to glyceraldehyde 3-phosphate (GAP). (b) The Wood-Ljungdahl pathway (WLP) converts two CO2 to acetyl-CoA. (c) The reductive glycine pathway (rGlyP) converts two CO2 to glycine. With one more CO2, glycine can be converted to serine and further to pyruvate and acetyl-CoA to yield more value-added products. (d) The serine cycle converts CO2 and 5,10-methylenetetrahydrofolate (5,10-CH2-THF), which can be derived from methane, methanol or formaldehyde, to acetyl-CoA. (e) The ribulose monophosphate cycle (RuMP) converts methane, methanol or formaldehyde to triose phosphate, which can feed into central carbon metabolism to produce cellular energy, biomass precursors, and bioproducts. Dark green boxes represent C1 substrates, the light green boxes represent potential products. Rate-limiting steps are underlined. ATP and redox cofactors are color-coded.
The WLP is an anoxic alternative to the CBB cycle for direct C2 synthesis (Figure 3b) [5]. In the WLP (also known as the reductive acetyl-CoA pathway), two CO2 molecules are reduced in parallel via the methyl and carbonyl branches to form acetyl-CoA. The methyl branch of the WLP is one-carbon metabolism that is conserved across the kingdoms of life [5]. Methyl transfer has been suggested as the rate-limiting step of the WLP [19]. ATP is used for activating C1 in the methyl branch and a rate-limiting factor in autotrophy even though ATP is regenerated during acetate production. Controlled cofeeding of glucose to supply ATP stimulates CO2 conversion into acetate [20]. In acetogens, Rnf and Nfn complexes contribute to energy conservation by chemiosmotic gradient buildup and electron bifurcation, respectively [21]. In Moorella thermoacetica, methylenetetrahydrofolate reductase (MTHFR), which converts methylene-THF to methyl-THF, is hypothesized to be electron bifurcating [22, 23]. Further studies into the energy-conserving mechanisms of the WLP would contribute to increasing the efficiency of input energy stored as energy-dense molecules [24]. The WLP serves as a steppingstone to converting CO2 and CO into advanced bioproducts by producing acetate that heterotrophic microbes can utilize as sole carbon and energy sources. Two-stage processes converting syngas containing CO2 and CO into acetate (e.g., using Clostridium aceticum and Moorella thermoacetica) and acetate into lipids (e.g., using Rhodotorula toruloides and Yarrowia lipolytica) have been demonstrated [20, 25, 26].
A partially reduced carbon formate is more water soluble compared to CO2 and CO, albeit toxic at moderate concentrations (~150 mM) [27]. Formate can be reduced through the WLP methyl branch and the reductive glycine pathway (rGlyP). The rGlyP converts formate and CO2 into glycine and with an additional formate into serine, which can be further converted into pyruvate (Figure 3c) [28]. In converting formate to glycine and serine, the rGlyP consumes ATP, NADPH, and NADH. While the reduction of 5,10-methenyl-THF to 5,10-methylene-THF is often rate-limiting in the rGlyP [28, 29], low ammonia concentrations has been shown to be a limiting factor in the rGlyP of Desulfovibrio desulfuricans [7]. The rGlyP is more ATP-efficient than the CBB cycle, it has proven efficient for C1 assimilation for cell growth and function in Pseudomonas putida and E. coli [29, 30].
Methane, methanol, and formaldehyde are challenging C1 substrates due to the gaseous nature of methane and the toxicity of formaldehyde. Life has still prevailed in evolving metabolic processes for detoxification and utilization of these substrates. Methane utilization in methanotrophs start by methyl transfer to C1 carriers or partial oxidation to methanol by using methane monooxygenase (MMO) [8, 31–33]. Mammals can dehydrogenate methanol to formaldehyde using alcohol dehydrogenase and further oxidize it to formate using aldehyde dehydrogenase [34]. Unicellular organisms like Pichia pastoris also use formaldehyde dehydrogenase to enable growth on methanol or formaldehyde. These relatively electron-rich carbons undergo both assimilative and dissimilative pathways for biomass production and energy generation [35]. To incorporate reduced C1 compounds into central carbon metabolism, type II methanotrophs use the serine cycle [36]. The serine cycle utilizes reduced C1 or formaldehyde and CO2 to form a C2 acetyl group (Figure 3d) [37, 38]. A synthetic serine cycle has been introduced in E. coli for methanol assimilation [38]. However, its high energy cost of 2 ATP per cycle is costlier methylotrophy than that using the ribulose monophosphate (RuMP) cycle [39].
The RuMP cycle (Figure 3e) is an energy efficient methanol assimilation pathway with a Gibbs free energy change (ΔG’°) of −89.9 kJ/mol [40]. The RuMP cycle is popular for synthetic methylotrophy in E. coli because of minimal enzyme expression requirements [39]. E. coli and other organisms have been successfully engineered with the RuMP cycle [39, 41, 42]. Keller and colleagues identified 3-hexulose-6-phosphate synthase (HPS), which condenses formaldehyde and ribulose-5-phosphate (Ru5P) into hexulose-6-phosphate (Hu6P), as the limiting reaction [39]. Sanford and Woolston recognized that methanol assimilation through the RuMP cycle is thermodynamically constrained by formaldehyde buildup [43]. One drawback of this pathway is that when the three-carbon product, dihydroxyacetone phosphate (DHAP), is converted to acetyl-CoA for use in lipogenesis, one of the carbons is lost as CO2. Thus, synthesizing products derived from C3 lower glycolytic intermediates is more desirable when using the RuMP cycle. The initial C1 compound and the desired final product should be paired using optimal pathways in terms of their carbon yield, ATP requirement, and heterologous gene expression.
Engineering Synthetic One-Carbon Metabolism
Integrating heterologous metabolic pathways into model organisms is a key objective of metabolic engineering. The benefit of using model organisms is that their well-characterized growth and metabolism makes them well suited for large-scale manufacturing [10]. By integrating natural one-carbon pathways into model organisms, researchers can get the best of both worlds. Chen and colleagues engineered a methylotrophic E. coli by introducing a modified RuMP cycle (Figure 4a) with three heterologous genes encoding methanol dehydrogenase (MDH), HPS, and 6-phospho-3-hexuloisomerase (PHI) [44]. To prevent formaldehyde toxicity, the engineered cells must keep the HPS reaction active and the RuMP cycle turning. The modified RuMP cycle minimizes the diversion of carbon flux away from HPS by using the Entner-Doudoroff (ED) pathway instead of PFK and fructose-1,6-bisphosphate aldolase (ALDO) for converting fructose-6-phosphate (F6P) into C3 lower glycolytic metabolites, one of which is the output of the RuMP cycle from three input C1 molecules. Using the ED pathway provides dual benefits: 1) the modified RuMP cycle produces pyruvate in fewer steps compared to the original RuMP cycle; and 2) the modified RuMP cycle does not require initial ATP investment [45]. By incorporating the modified RuMP cycle, Reiter and colleagues achieved synthetic methylotrophy in E. coli with a 4.3-hour doubling time and maximum culture density of 100 OD600 in a fed-batch reactor that continuously fed and maintained methanol at >500 mM [46].
Figure 4. Synthetic C1 assimilation pathways integrated into host metabolism.
(a) A modified RuMP cycle introduced in E. coli leverages the ED pathway to supply C3 intermediates without ATP expenditure [44]. (b) Synthetic EuMP cycle converts formaldehyde to E4P with the introduction of EPS, LerI, and DerI enzymes. The cycle is completed by central carbon reactions. Adapted from [47]. (c) The synthetic homoserine cycle condenses formaldehyde with pyruvate to produce HOB and eventually acetyl-CoA. Adapted from [48]. (d) The HOB cycle is a synthetic alternative to one-carbon metabolism. Adapted from [49]. (e) Phosphate dependent formate to formaldehyde conversion can feed into formyl-CoA elongation reactions (FORCE) for synthesis of multicarbon units [50, 54]. Adapted from [54]. Dark green boxes represent C1 substrates, the light green boxes represent potential products. ATP and redox cofactors are color-coded. Segments of the pathways that are either modified, introduced in the organism or new-to-nature to create these cycles are highlighted in orange.
Designing novel pathways augments the toolset for C1 utilization. Wu and colleague designed the erythrulose monophosphate (EuMP) cycle for formaldehyde assimilation. Formaldehyde is condensed with DHAP to produce erythrulose-1-phosphate (Eu1P) and the product after three turns of the cycle is GAP (Figure 4b) [47]. The EuMP cycle is energetically efficient with just one ATP consumed per GAP produced. While it is just as energetically efficient as the RuMP cycle, the EuMP cycle includes more enzymatic steps.
The use of orthogonal molecules mitigates the potential issue of substrate competition between native and engineered pathways. The synthetic homoserine cycle for methanol assimilation condenses formaldehyde with pyruvate to produce 4-hydroxy-2-oxobutanoic acid (HOB), which can be converted to homoserine (Figure 4c) [48]. Similarly, a synthetic 4-hydroxy-2-oxobutanoic acid (HOB)-dependent one-carbon metabolism has been introduced in E. coli as an alternative to the serine cycle (Figure 4d) [49]. This pathway relies on two novel reactions: the transamination of L-homoserine and transfer of one carbon unit from HOB to THF with release of pyruvate. The strain assimilates CO2 without relying on endogenous metabolism. Chou and colleagues designed an orthogonal metabolic framework in E. coli for creating multi-carbon products directly from C1 substrates using formyl-CoA elongation reactions (FORCE) so that the assimilation does not compete with the host metabolism [50]. Intermediates from these pathways can be readily transformed to several value-added products (2-hydroxy acids, aldoses, diols, polyols, carboxylic acids and alcohols) as well as biomass precursors. Thermodynamic and stoichiometric analyses identified ATP consumption, the NADH/NAD+ ratio, and formate concentration as major constraints to the max-min driving force (MDF) of the FORCE pathway [51]. The FORCE pathway was introduced in E. coli and tested in a co-culture system, in which one strain could convert formate, methanol, and formaldehyde to glycolate while the other could consume glycolate for growth.
Engineering new enzymes benefits C1 utilization on an individual reaction level and helps conceive new pathways. The condensation of formyl-CoA and formaldehyde is a major bottleneck in the FORCE pathway [50, 52]. The oxalyl-CoA decarboxylase from Methylobacterium extorquens was mutated to a glycolyl-CoA synthase (MeOXC4), which allows for C1-C1 condensation of formyl-CoA and formaldehyde. MeOXC4 had 40-fold more catalytic efficiency compared to all other known natural and synthetic enzymes for C1-C1 condensation at the time [53]. Since then, the variants of 2-hydroxy-CoA synthase (HACS) with improved activities have been identified [52]. Nattermann and colleagues also introduced a two-step cascade for phosphate-dependent conversion of formate to formaldehyde [54]. The new-to-nature enzyme formyl phosphate reductase (FPR) boasts a 300-fold shift in formyl phosphate specificity. Combining FPR with the FORCE pathway [50] allows C2 production with formate as the sole carbon source (Figure 4e).
Powering Metabolism to Attain the Circular Economy
A common requirement for operating one-carbon metabolism is energy in the form of reducing power and high-energy phosphates. For example, acetogenesis requires 2 CO2 and 8 electrons with the redox potential of CO2/acetate of −0.28 V versus standard hydrogen electrode (SHE) [55]. In biological systems, ATP is required for driving thermodynamically unfavorable reactions and running cellular housekeeping processes. Substrate cofeeding combined with systems and synthetic biology tools increases the efficiency of C1 conversion by better affording cellular energy. Furthermore, electrochemical cells and light-harvesting materials provide alternative means to power bioenergetic pathways. Utilization of diversified energy sources will engender a resilient circular bioeconomy.
Balanced Cofactor Generation Powers Metabolism
Efficient bioproduct synthesis using C1 compounds necessitates a balanced supply of carbon backbones, ATP, and reducing equivalents (Figure 5a). Canonical redox pairs are ubiquitously employed across metabolism (Box 1). Healthy metabolic function hinges on their availability [56]. Synthetic biology efforts have sought to improve and balance cofactor generation [57, 58], favor one cofactor over another [59], and introduce non-canonical redox cofactors to minimize loss of electrons to competing pathways [56]. One approach to overcoming imbalanced cofactor production is substrate cofeeding, which activates different metabolic pathways without genetic engineering [60]. Feeding C1 substrates with sugars such as glucose and fructose has been shown to improve energy efficiency and carbon yield. Doping gas-fermenting M. thermoacetica with glucose streamlines ATP generation to accelerate cellular metabolism and CO2 reduction [20]. Limiting glucose supply limits its accumulation in the culture and overcomes catabolite repression in the mixed substrate cofeeding while activating glycolytic ATP generation. A mixture of CO2, CO, and H2 can be fed to acetogens to power the WLP. In the WLP, CO2 and CO provide carbon backbone whereas the oxidation of CO and H2 provide reducing equivalents to create NAD(P)H and reduced ferredoxin. Mixotrophic growth on C1 compounds, sugars, and other nutrients promotes fast metabolism and bioproduct synthesis [61–63]. Another approach to overcome redox cofactor limitations is to utilize alternative cofactors, which has been implemented in E. coli to support cell growth and channel reducing power from glucose to levodione through nicotinamide mononucleotide (NMN) rather than NADPH [64]. The cofactor specificity of enzymes for NADH and NADPH may be swapped. The mevalonate pathway for isoprenoid synthesis uses either NADPH or NADH depending on the organism. In Halomonas bluephagenis, NADH-dependent strains produced more mevalonate because NADH is more abundant in the cell [65]. In engineered E. coli methylotrophs, the MDH reaction which converts methanol to formaldehyde is inhibited by high NADH concentrations. To overcome this, researchers have sought to couple methanol oxidation to an artificial NADH consumption pathway [66]. These synthetic biology approaches to boost the availability and efficient use of redox cofactors help address challenges in C1 substrate utilization.
Figure 5. Chemical, electric, and solar energy powers metabolism.
(a) Balanced supply of carbons, reducing power, and ATP via metabolic engineering and substrate cofeeding accelerates C1 utilization. (b) Electrochemistry-powered bioenergetic pathway. Carrier molecules (e.g., H2 and syngas) transfer reducing power from cathode to bacteria and support C1 conversion and autotrophy. (c) Light-powered bioenergetic pathway. Light-harvesting materials (e.g., quantum dots) harness solar power and transfer reducing equivalent to microbes through cell-material interface, creating an artificial photosynthetic system.
Box 1. Efficient energy utilization using disparate redox cofactors.
Redox cofactors mediate electron transfers and couple anabolic and catabolic processes in metabolism. Each redox pair has a unique redox potential. Enzymatic steps in C1 assimilation pathways utilize different redox cofactors (Figure 3). Nicotinamide cofactors (NAD and NADP and their reduced equivalents NADH and NADPH) both have reduction potential E0’=–320mV however their redox potentials differ due to the availability of their reduced and oxidized pools [56]. The NAD/NADH ratio is kept very high by shuttling electrons to the electron transport chain and is used in oxidation reactions. The NADP/NADPH ratio is kept low due to the role of NADPH as a reducing power in biosynthesis. NADPH is produced by the oxidative pentose phosphate pathway and malic enzyme as well as one-carbon metabolism and isocitrate dehydrogenase in some organisms. Ferredoxins, on the other hand, are iron-sulfur proteins with reduction potential E0’~–420mV [87, 88]. In contrast to nicotinamide cofactors which mediated two electron transfers at time, ferredoxins typically mediate a single electron transfer. Cells have ferredoxins with different Fe-S clusters, each tuned for specific reactions. We have shown the most reported cofactors for each reaction (Figure 3), however different species may possess different enzymes to carry out the same reaction using different cofactors. A more detailed view of these pathways can be found on MetaCyc (MetaCyc.org), where for example the methylene tetrahydrofolate reaction (MTHFR) in the WLP shows three possible cofactor pairs for the reduction reaction and different possible enzymes. Interestingly, MTHFR reaction in M. thermoacetica is thought to operate electron bifurcation to minimize overpotential. Electron bifurcation transfers electrons from one source to two disparate electron carriers, one with more negative reduction potential for overcoming “difficult” energy barriers (e.g., in CO2 reduction) and the other with less negative reduction potential to limit wasting free energy [88]. Incorporating electron bifurcation in anaerobic digestion has been shown to increase ATP generation and methane production under substrate limitation [89]. Introducing a flavin-based electron bifurcation in E. coli led up to a 4-fold increase in H2 and succinate production, 7-fold increase in ATP accumulation and superior cell-fitness [90].
Electrochemistry Powers Metabolism
In electrochemistry-assisted microbial CO2 fixation, cathodes power CO2 fixation by transferring reducing equivalents such as H2 to cells (Figure 5b). A cobalt-phosphorus alloy cathode has been used to produce H2 to enhance CO2 fixation process of Ralstonia eutropha [67]. Deutzmann and colleagues designed a hybrid biochemical CO2 fixation system including a biocathode consisting of a graphite electrode and Desulfopila corrodens strain IS4 [68]. H2 production speed on the biocathode exceeds that of platinum electrode, a widely used inorganic catalyst. The feasibility of the biocathode-driven CO2 fixation has been demonstrated using a methanogen Methanococcus maripaludis and an acetogen Acetobacterium woodii for methane and acetate production. A photovoltaic device has been used to electrochemically reduce CO2 and H2O into syngas which was further transformed into butanol and hexanol using Clostridium autoethanogenum [69]. Remarkably, this hybrid process featured faradaic efficiency as high as 80–100% and a lower energy cost compared to a chemical synthesis route.
Electrochemistry-assisted bacterial CO2 fixation has lower energy costs and higher efficiency than photosynthesis, but maximum CO2 conversion efficiency is limited by the low solubility of electron mediators (e.g., H2). To overcome the obstacle, perfluorocarbon (PFC) nano-emulsion has been used as a H2 carrier to promote acetate production by Sporomusa ovata [70]. Due to PFC’s excellent H2 solubility, addition of nano-emulsion increases the throughput of CO2 reduction into acetate by as much as 190%. Claassens and colleagues found that formate and methanol can work as electron mediators to supply reducing power from electrodes to microbes at a high energetic efficiency [71]. Despite the simple nature of acetate, electrochemistry-driven CO2 reduction into C2 products is beneficial since heterotrophic microbes can utilize the C2 compounds as the sole carbon and energy source [72].
Besides providing reducing power, electrochemistry enables fine tuning of microenvironments. Liu and colleagues utilized Pt-coated Si wire array to achieve CO2-to-acetate transformation of a strict anaerobe S. ovata under aerobic conditions [73]. With solar energy input, Pt catalyzes O2 reduction by photocurrent in the local environment, leading to microscopic anoxic conditions to support the S. ovata culture. Similarly, electrochemical O2 microscopic gradients have been used to accommodate N2-fixing bacteria [74] for accelerated nitrogen assimilation. A unique benefit of introducing electrochemistry to bioprocesses is the control it offers. By adjusting applied voltage and electrode morphology, energy can be efficiently delivered to cells. Machine learning has proven useful in tuning these variables to create suitable microenvironments for microbial C1 conversion to afford promising avenues for biomass and bioproduct synthesis [75].
Light-Harvesting Material Powers Metabolism
Light-harvesting materials can transduce solar energy into reducing power for microbes (Figure 5c). Material-bacteria interfaces thus give rise to artificial photosynthetic systems. One commonly used light-harvesting material is Si wire array. A Si-nanowire-bacteria hybrid was the first example of direct interface between CO2-reducing S. ovata and semiconductors allowing microbial photoelectrosynthesis for CO2-to-acetate transformation [73]. Adding on an N2-fixing bacterium Rhodopseudomonas palustris to the nanowire-bacteria system, the co-culture system achieved a solar-to-chemical efficiency of 0.51% for nitrogenous biomass [76]. A hybrid of quantum dots and bacteria also creates an artificial photosynthesis system [77]. CdS quantum dots can capture sunlight and produce reducing equivalents to operate the WLP in M. thermoacetica. CdTe quantum dots have been interfaced as light harvesters on Xanthobacter autotrophicus for simultaneous CO2 and N2 fixation [78]. High photophysical charge-transfer kinetics between quantum dots and microbes enables CO2 fixation at an internal quantum efficiency of ~47%, close to the biochemical limits at 46.1%. Semiconductors La- and Rh-doped SrTiO3 and Mo-doped BiVO4 have been used as a photocatalyst sheet for absorbing solar energy for water splitting, and providing S. ovata with H2 for the CO2-to-acetate conversion [79]. An organic semiconductor (e.g., a heterojunction of perylene diimide derivative and poly(fluorene-co-phenylene)) has been used to generate photoexcited electrons for nonphotosynthetic bacteria M. thermoacetica to reduce CO2 into acetate [80]. Thus, light-harvesting materials expand the repertoire of renewable energy sources for microbial C1 conversion.
Toward hybrid powering of metabolism
The ability of microbes to harness multiple energy sources promotes bioproduct synthesis, rain or shine. The versatile energy utilization can be accomplished on an individual culture level, in which an isogenic or heterogeneous population of cells have access to chemical, electrical, and light energy, or in distributed cultures that work in concert to upcycle C1 compounds. The hybrid powering strategy may combine the benefits of nutrient cofeeding that activates multiple parts of metabolism for advanced bioproduct synthesis, electrochemistry that confers excellent controllability, and abiological material that harvests and transduces plentiful solar energy. We envision interdisciplinary projects coordinating to construct a “power grid” for bioeconomy.
Concluding Remarks: Bringing Elements to Life
Engineering and powering metabolism through cofactor balancing and green energy provide a blueprint for repurposing of C1 compounds to value-added products. Central carbon metabolism acts as a rendezvous for one-carbon metabolism and biosynthetic metabolism. Matchmaking between C1 substrates and advanced bioproducts requires contiguous enzymatic reaction steps that bridge them and energy metabolism that powers the assimilative and biosynthetic pathways. Thus, systems-level biochemical and energetic considerations help conceive scientifically and commercially viable C1-bioconversion processes.
Integration of analytical, biochemical, and computational tools provides a holistic understanding of one-carbon metabolism from thermodynamic and kinetic perspectives (see Outstanding Questions). In heterotrophic organisms, mass spectrometry, 13C stable isotope tracing, and mathematical optimization have imparted flux information (i.e., rates at which cells operate metabolic pathways) based on steady-state isotope labeling patterns in metabolites that imprint metabolic activities. In autotrophs, 13C tracing from C1 compounds is more difficult as the steady-state 13C-labeling state comprises uniformly labeled metabolites, and thus dynamic labeling strategies are preferred [81–83]. A quantitative understanding of one-carbon metabolism flux and regulation will engender more rational metabolic engineering and synthetic biology endeavors for C1 upcycling.
Outstanding Questions.
Cells are made of several major elements including carbon and nitrogen. Thus, carbon and nitrogen metabolism must be coordinated for efficient biomass and bioproduct synthesis. How can we better understand the regulation and coordination of carbon and nitrogen metabolism? How can the resulting knowledge help repurpose nitrogen waste simultaneously with carbon waste?
Problems at hand require interdisciplinary approaches. How can biochemical, genetic, and computational tools be integrated to facilitate efficient energy transduction, biosynthesis, and their coupling?
What are technoeconomically feasible C1-to-product conversion processes?
How can we translate laboratory knowledge to industrial scale bioprocesses to engender a positive global impact?
From a technoeconomic perspective, additional considerations of scale-up, market sizing, and process design are needed. Only when the biological process yields cost-effective products at scale compared to existing chemical processes would value be created. For example, LanzaTech has successfully commercialized a gas-fermentation process using syngas from steel mills to produce ethanol and acetone using acetogenic Clostridia [84]. Others have worked towards commercial syngas fermentation; however, challenges related to scale-up, gas-to-liquid mass transfer, and costs linger [85]. While powering one-carbon metabolism to convert C1 substrates into C2 and C3 intermediate products acts as a store of value, attaining C4+ products can drive value creation. Commercialization of C1-derived jet fuels and advanced bioproducts such as polyketides and terpenoids would yield environmentally and economically productive outcomes.
In addition to carbon, nitrogen constitutes a large portion of the biogeochemical cycle. Nitrogen-fixing bacteria as well as the Haber-Bosch process produce ammonia from N2. Nitrate and nitrite from runoff wastewater may also be converted to ammonia using electrochemical processes [86]. Integrative research into carbon and nitrogen metabolism will bring us closer to a circular economy (see Outstanding Questions).
ACKNOWLEDGEMENTS
The authors would like to thank the members of the Park and Liu labs for helpful discussion. This work was supported by the National Institute of General Medical Sciences of the National Institutes of Health under Award Number R35GM143127 (J.O.P.), the Department of Energy under Award Number DE-SC0024251 (J.O.P.), and the Center for Clean Technology Fellowship (R.C.L.).
Footnotes
Declaration of Interests
The authors declare no competing interests.
REFERENCES
- 1.Jin S, et al. (2021) Advances and challenges for the electrochemical reduction of CO2 to CO: from fundamentals to industrialization. Angewandte Chemie 133, 20795–20816 [DOI] [PubMed] [Google Scholar]
- 2.Han N, et al. (2020) Promises of main group metal–based nanostructured materials for electrochemical CO2 reduction to formate. Advanced Energy Materials 10, 1902338 [Google Scholar]
- 3.Wang H, et al. (2023) CO2 electrolysis towards acetate: a review. Current Opinion in Electrochemistry, 101253 [Google Scholar]
- 4.Overa S, et al. (2022) Electrochemical approaches for CO2 conversion to chemicals: a journey toward practical applications. Accounts of Chemical Research 55, 638–648 [DOI] [PubMed] [Google Scholar]
- 5.Ragsdale SW and Pierce E (2008) Acetogenesis and the Wood–Ljungdahl pathway of CO2 fixation. Biochimica et Biophysica Acta (BBA)-Proteins and Proteomics 1784, 1873–1898 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Asplund-Samuelsson J and Hudson EP (2021) Wide range of metabolic adaptations to the acquisition of the Calvin cycle revealed by comparison of microbial genomes. PLOS Computational Biology 17, e1008742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Sánchez-Andrea I, et al. (2020) The reductive glycine pathway allows autotrophic growth of Desulfovibrio desulfuricans. Nature Communications 11, 5090. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Bennett RK, et al. (2018) Engineering the bioconversion of methane and methanol to fuels and chemicals in native and synthetic methylotrophs. Current opinion in biotechnology 50, 81–93 [DOI] [PubMed] [Google Scholar]
- 9.Law RC, et al. (2022) Integrative metabolic flux analysis reveals an indispensable dimension of phenotypes. Current opinion in biotechnology 75, 102701. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Nurwono G, et al. (2023) Sustainable metabolic engineering requires a perfect trifecta. Current Opinion in Biotechnology 83, 102983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Mostaghimi AHB, et al. (2020) A review on electrocatalytic oxidation of methane to oxygenates. Journal of Materials Chemistry A 8, 15575–15590 [Google Scholar]
- 12.Schuchmann K and Müller V (2014) Autotrophy at the thermodynamic limit of life: a model for energy conservation in acetogenic bacteria. Nature Reviews Microbiology 12, 809–821 [DOI] [PubMed] [Google Scholar]
- 13.Erşan S and Park JO (2020) Light-independent biological conversion of CO2. Joule 4, 2047–2051 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Sharkey TD (2019) Discovery of the canonical Calvin–Benson cycle. Photosynthesis Research 140, 235–252 [DOI] [PubMed] [Google Scholar]
- 15.Raines CA (2022) Improving plant productivity by re‐tuning the regeneration of RuBP in the Calvin–Benson–Bassham cycle. New Phytologist 236, 350–356 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Cordero PR, et al. (2019) Atmospheric carbon monoxide oxidation is a widespread mechanism supporting microbial survival. The ISME journal 13, 2868–2881 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Antonovsky N, et al. (2016) Sugar synthesis from CO2 in Escherichia coli. Cell 166, 115–125 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Yu J, et al. (2023) Increase CO2 recycling of Escherichia coli containing CBB genes by enhancing solubility of multiple expressed proteins from an operon through temperature reduction. Microbiology Spectrum 11, e02560–02523 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Seravalli J, et al. (2002) Rapid kinetic studies of Acetyl-CoA synthesis: evidence supporting the catalytic intermediacy of a paramagnetic NiFeC species in the autotrophic wood− ljungdahl pathway. Biochemistry 41, 1807–1819 [DOI] [PubMed] [Google Scholar]
- 20.Park JO, et al. (2019) Synergistic substrate cofeeding stimulates reductive metabolism. Nature Metabolism 1, 643–651 [DOI] [PubMed] [Google Scholar]
- 21.Marcellin E, et al. (2016) Low carbon fuels and commodity chemicals from waste gases–systematic approach to understand energy metabolism in a model acetogen. Green Chemistry 18, 3020–3028 [Google Scholar]
- 22.Wang S, et al. (2013) A reversible electron-bifurcating ferredoxin-and NAD-dependent [FeFe]-hydrogenase (HydABC) in Moorella thermoacetica. Journal of bacteriology 195, 1267–1275 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Kremp F and Müller V (2021) Methanol and methyl group conversion in acetogenic bacteria: biochemistry, physiology and application. FEMS Microbiology Reviews 45, fuaa040. [DOI] [PubMed] [Google Scholar]
- 24.Latif H, et al. (2014) Trash to treasure: production of biofuels and commodity chemicals via syngas fermenting microorganisms. Current Opinion in Biotechnology 27, 79–87 [DOI] [PubMed] [Google Scholar]
- 25.Robles-Iglesias R, et al. (2023) Sequential bioconversion of C1-gases (CO, CO2, syngas) into lipids, through the carboxylic acid platform, with Clostridium aceticum and Rhodosporidium toruloides. Journal of Environmental Management 347, 119097. [DOI] [PubMed] [Google Scholar]
- 26.Hu P, et al. (2016) Integrated bioprocess for conversion of gaseous substrates to liquids. Proceedings of the National Academy of Sciences 113, 3773–3778 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Sharak Genthner B and Bryant M (1987) Additional characteristics of one-carbon-compound utilization by Eubacterium limosum and Acetobacterium woodii. Applied and Environmental Microbiology 53, 471–476 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Kim S, et al. (2020) Growth of E. coli on formate and methanol via the reductive glycine pathway. Nature chemical biology 16, 538–545 [DOI] [PubMed] [Google Scholar]
- 29.Yishai O, et al. (2018) In vivo assimilation of one-carbon via a synthetic reductive glycine pathway in Escherichia coli. ACS synthetic biology 7, 2023–2028 [DOI] [PubMed] [Google Scholar]
- 30.Bruinsma L, et al. (2023) Paving the way for synthetic C1-metabolism in Pseudomonas putida through the reductive glycine pathway. Metabolic Engineering 76, 215–224 [DOI] [PubMed] [Google Scholar]
- 31.Chadwick GL, et al. (2022) Comparative genomics reveals electron transfer and syntrophic mechanisms differentiating methanotrophic and methanogenic archaea. PLoS biology 20, e3001508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Evans PN, et al. (2019) An evolving view of methane metabolism in the Archaea. Nature Reviews Microbiology 17, 219–232 [DOI] [PubMed] [Google Scholar]
- 33.Leu AO, et al. (2020) Anaerobic methane oxidation coupled to manganese reduction by members of the Methanoperedenaceae. The ISME Journal 14, 1030–1041 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Pietzke M, et al. (2020) Amino acid dependent formaldehyde metabolism in mammals. Communications Chemistry 3, 78. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Berrios J, et al. (2022) Role of dissimilative pathway of Komagataella phaffii (Pichia pastoris): formaldehyde toxicity and energy metabolism. Microorganisms 10, 1466. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Hanson RS and Hanson TE (1996) Methanotrophic bacteria. Microbiological reviews 60, 439–471 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Klein VJ, et al. (2022) Unravelling formaldehyde metabolism in bacteria: road towards synthetic methylotrophy. Microorganisms 10, 220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Yu H and Liao JC (2018) A modified serine cycle in Escherichia coli coverts methanol and CO2 to two-carbon compounds. Nature communications 9, 3992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Keller P, et al. (2020) Methanol-dependent Escherichia coli strains with a complete ribulose monophosphate cycle. Nature communications 11, 5403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Flamholz A, et al. (2012) eQuilibrator—the biochemical thermodynamics calculator. Nucleic acids research 40, D770–D775 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Müller JE, et al. (2015) Engineering Escherichia coli for methanol conversion. Metabolic engineering 28, 190–201 [DOI] [PubMed] [Google Scholar]
- 42.Whitaker WB, et al. (2017) Engineering the biological conversion of methanol to specialty chemicals in Escherichia coli. Metabolic engineering 39, 49–59 [DOI] [PubMed] [Google Scholar]
- 43.Sanford PA and Woolston BM (2022) Synthetic or natural? Metabolic engineering for assimilation and valorization of methanol. Current Opinion in Biotechnology 74, 171–179 [DOI] [PubMed] [Google Scholar]
- 44.Chen FY-H, et al. (2020) Converting Escherichia coli to a synthetic methylotroph growing solely on methanol. Cell 182, 933–946. e914 [DOI] [PubMed] [Google Scholar]
- 45.Law RC, et al. (2024) A parallel glycolysis provides a selective advantage through rapid growth acceleration. Nature Chemical Biology 20, 314–322 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Reiter MA, et al. (2024) A synthetic methylotrophic Escherichia coli as a chassis for bioproduction from methanol. Nature Catalysis, 1–14 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Wu T, et al. (2023) Engineering a synthetic energy-efficient formaldehyde assimilation cycle in Escherichia coli. Nature Communications 14, 8490. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.He H, et al. (2020) An optimized methanol assimilation pathway relying on promiscuous formaldehyde-condensing aldolases in E. coli. Metabolic engineering 60, 1–13 [DOI] [PubMed] [Google Scholar]
- 49.Bouzon M, et al. (2017) A synthetic alternative to canonical one-carbon metabolism. ACS synthetic biology 6, 1520–1533 [DOI] [PubMed] [Google Scholar]
- 50.Chou A, et al. (2021) An orthogonal metabolic framework for one-carbon utilization. Nature Metabolism 3, 1385–1399 [DOI] [PubMed] [Google Scholar]
- 51.Noor E, et al. (2014) Pathway thermodynamics highlights kinetic obstacles in central metabolism. PLoS computational biology 10, e1003483. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Lee SH, et al. (2023) Identification of 2-Hydroxyacyl-CoA synthases with high acyloin condensation activity for orthogonal one-carbon bioconversion. ACS Catalysis 13, 12007–12020 [Google Scholar]
- 53.Nattermann M, et al. (2021) Engineering a highly efficient carboligase for synthetic one-carbon metabolism. ACS catalysis 11, 5396–5404 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Nattermann M, et al. (2023) Engineering a new-to-nature cascade for phosphate-dependent formate to formaldehyde conversion in vitro and in vivo. Nature Communications 14, 2682. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Rabaey K and Rozendal RA (2010) Microbial electrosynthesis — revisiting the electrical route for microbial production. Nature Reviews Microbiology 8, 706–716 [DOI] [PubMed] [Google Scholar]
- 56.Weusthuis RA, et al. (2020) Applying non-canonical redox cofactors in fermentation processes. Iscience 23 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Wang Y, et al. (2022) Reassessing acetyl‐CoA supply and NADPH availability for mevalonate biosynthesis from glycerol in Escherichia coli. Biotechnology and Bioengineering 119, 2868–2877 [DOI] [PubMed] [Google Scholar]
- 58.Jaroensuk J, et al. (2024) A versatile in situ cofactor enhancing system for meeting cellular demands for engineered metabolic pathways. Journal of Biological Chemistry 300 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Ma SM, et al. (2011) Optimization of a heterologous mevalonate pathway through the use of variant HMG-CoA reductases. Metabolic engineering 13, 588–597 [DOI] [PubMed] [Google Scholar]
- 60.Liu N, et al. (2020) Mixed carbon substrates: a necessary nuisance or a missed opportunity? Current Opinion in Biotechnology 62, 15–21 [DOI] [PubMed] [Google Scholar]
- 61.Maru BT, et al. (2018) Fixation of CO2 and CO on a diverse range of carbohydrates using anaerobic, non-photosynthetic mixotrophy. FEMS microbiology letters 365, fny039. [DOI] [PubMed] [Google Scholar]
- 62.Bae J, et al. (2022) Valorization of C1 gases to value-added chemicals using acetogenic biocatalysts. Chemical Engineering Journal 428, 131325 [Google Scholar]
- 63.Kang Y, et al. (2022) A two-stage process for the autotrophic and mixotrophic conversion of C1 gases into bacterial cellulose. Bioresource Technology 361, 127711. [DOI] [PubMed] [Google Scholar]
- 64.Black WB, et al. (2020) Engineering a nicotinamide mononucleotide redox cofactor system for biocatalysis. Nature chemical biology 16, 87–94 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Zhang J, et al. (2023) Metabolic engineering of Halomonas bluephagenesis for high-level mevalonate production from glucose and acetate mixture. Metabolic Engineering 79, 203–213 [DOI] [PubMed] [Google Scholar]
- 66.Price JV, et al. (2016) Scaffoldless engineered enzyme assembly for enhanced methanol utilization. Proceedings of the National Academy of Sciences 113, 12691–12696 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Liu C, et al. (2016) Water splitting–biosynthetic system with CO2 reduction efficiencies exceeding photosynthesis. Science 352, 1210–1213 [DOI] [PubMed] [Google Scholar]
- 68.Deutzmann JS and Spormann AM (2017) Enhanced microbial electrosynthesis by using defined co-cultures. The ISME Journal 11, 704–714 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Haas T, et al. (2018) Technical photosynthesis involving CO2 electrolysis and fermentation. Nature Catalysis 1, 32–39 [Google Scholar]
- 70.Rodrigues RM, et al. (2019) Perfluorocarbon nanoemulsion promotes the delivery of reducing equivalents for electricity-driven microbial CO2 reduction. Nature Catalysis 2, 407–414 [Google Scholar]
- 71.Claassens NJ, et al. (2019) Making quantitative sense of electromicrobial production. Nature Catalysis 2, 437–447 [Google Scholar]
- 72.Zhang P, et al. (2022) Chem-bio interface design for rapid conversion of CO2 to bioplastics in an integrated system. Chem 8, 3363–3381 [Google Scholar]
- 73.Liu C, et al. (2015) Nanowire–Bacteria Hybrids for Unassisted Solar Carbon Dioxide Fixation to Value-Added Chemicals. Nano Letters 15, 3634–3639 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Lu S, et al. (2020) Electricity-powered artificial root nodule. Nature Communications 11, 1505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Chen Y, et al. (2022) Machine learning–based inverse design for electrochemically controlled microscopic gradients of O2 and H2O2. Proceedings of the National Academy of Sciences 119, e2206321119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Cestellos-Blanco S, et al. (2022) Photosynthetic biohybrid coculture for tandem and tunable CO2 and N2 fixation. Proceedings of the National Academy of Sciences 119, e2122364119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Sakimoto KK, et al. (2016) Self-photosensitization of nonphotosynthetic bacteria for solar-to-chemical production. Science 351, 74–77 [DOI] [PubMed] [Google Scholar]
- 78.Guan X, et al. (2022) Maximizing light-driven CO2 and N2 fixation efficiency in quantum dot–bacteria hybrids. Nature Catalysis 5, 1019–1029 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Wang Q, et al. (2022) Bacteria–photocatalyst sheet for sustainable carbon dioxide utilization. Nature Catalysis 5, 633–641 [Google Scholar]
- 80.Gai P, et al. (2020) Solar-Powered Organic Semiconductor–Bacteria Biohybrids for CO2 Reduction into Acetic Acid. Angewandte Chemie International Edition 59, 7224–7229 [DOI] [PubMed] [Google Scholar]
- 81.Hoyt KO and Woolston BM (2022) Adapting isotopic tracer and metabolic flux analysis approaches to study C1 metabolism. Current Opinion in Biotechnology 75, 102695. [DOI] [PubMed] [Google Scholar]
- 82.Rahim M, et al. (2022) INCA 2.0: A tool for integrated, dynamic modeling of NMR-and MS-based isotopomer measurements and rigorous metabolic flux analysis. Metabolic engineering 69, 275–285 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Abernathy MH, et al. (2017) Deciphering cyanobacterial phenotypes for fast photoautotrophic growth via isotopically nonstationary metabolic flux analysis. Biotechnology for biofuels 10, 1–13 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Tschaplinski TJ, et al. (2019) Development of a Sustainable Green Chemistry Platform for Production of Acetone and Downstream Drop-in Fuel and Commodity Products Directly From Biomass Syngas Via a Novel Energy Conserving Route in Engineered Acetogenic Bacteria. Oak Ridge National Lab.(ORNL), Oak Ridge, TN (United States) [Google Scholar]
- 85.Owoade A, et al. (2023) Progress and development of syngas fermentation processes toward commercial bioethanol production. Biofuels, Bioproducts and Biorefining 17, 1328–1342 [Google Scholar]
- 86.Murphy E, et al. (2022) Highly durable and selective Fe-and Mo-based atomically dispersed electrocatalysts for nitrate reduction to ammonia via distinct and synergized NO2–pathways. ACS Catalysis 12, 6651–6662 [Google Scholar]
- 87.Maiocco SJ, et al. (2019) Parsing redox potentials of five ferredoxins found within Thermotoga maritima. Protein Science 28, 257–266 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Buckel W and Thauer RK (2018) Flavin-based electron bifurcation, ferredoxin, flavodoxin, and anaerobic respiration with protons (Ech) or NAD+ (Rnf) as electron acceptors: a historical review. Frontiers in microbiology 9, 401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Zhang T, et al. (2023) Electronic bifurcation: a new perspective on Fe bio-utilization in anaerobic digestion of lactate. Environmental Science & Technology 57, 10448–10457 [DOI] [PubMed] [Google Scholar]
- 90.Intasian P, et al. (2023) Empowering extra fuel supply in E. coli by electron bifurcation for robust H2, ATP and succinate production. doi: 10.26434/chemrxiv-2023-fz0lz [DOI]
- 91.Silva CS, et al. (2015) Exergy efficiency of plant photosynthesis. Chemical engineering science 130, 151–171 [Google Scholar]
- 92.Bassham J and Krause G (1969) Free energy changes and metabolic regulation in steady-state photosynthetic carbon reduction. Biochimica et Biophysica Acta (BBA)-Bioenergetics 189, 207–221 [DOI] [PubMed] [Google Scholar]
- 93.Anthony C (2011) How half a century of research was required to understand bacterial growth on C1 and C2 compounds; the story of the serine cycle and the ethylmalonyl-CoA pathway. Science progress 94, 109–137 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Dick JM and Shock EL (2011) Calculation of the relative chemical stabilities of proteins as a function of temperature and redox chemistry in a hot spring. PLoS One 6, e22782. [DOI] [PMC free article] [PubMed] [Google Scholar]