Microbial C1-Fixation Dr. Simon K.-M. R. Rittmann Ferry, 2010 Global carbon cycle The global carbon cycle. Aerobic O2-requiring conversions are shown in the top panel and anaerobic conversions in the bottom panel. Step 1: Fixation of CO2 into organic matter. Step 2: Decomposition of organic matter to CO2 by O2-requiring microbes. Step 3: Deposition of organic matter into anaerobic environments. Step 4: Decomposition of complex biomass by fermentative microbes. Step 5: Conversion of volatile fatty acids by obligate H2- and formateproducing syntrophic acetogens. Step 6: H2- and formate-dependent reduction of CO2 to CH4 and conversion of the methyl group of acetate to CH4 by methanogens. Step 7: Anaerobic oxidation of CH4. Step 8: Diffusion of CH4 into aerobic zones. Step 9: Aerobic oxidation of CH4 by O2requiring methylotrophs. 2 (Dissolved) carbon dioxide (CO2), also in the form of bicarbonate (HCO3 –) Carbon monoxide (CO) Formate (HCOO–) Methane (CH4) Methanol (CH3OH) Mono-, di-, trimethylamine Methanethiole (CH3SH) C1 species 3 CO2 equilibrium in aqueous solution Wang et al., 2010 4 Currently six CO2 fixation pathways are known: Calvin–Benson–Bassham cycle reductive tricarboxylic acid cycle reductive acetyl CoA pathway – Wood-Ljungdahl pathway 3-hydroxypropionate bicycle 3 hydroxypropionate/4 hydroxybutyrate cycle dicarboxylate/4 hydroxybutyrate cycle (reductive hexulose-phosphate pathway) proposed [reductive glycine pathway] metagenome [Crotonyl-CoA/ethylmalonyl-CoA/hydroxybutyryl-CoA cycle] synthetic CO2 fixation pathways 5  In the Calvin–Benson–Bassham cycle, which was discovered about 50 years ago, CO2 reacts with the five-carbon sugar ribulose 1,5-bisphosphate to yield two carboxylic acids, 3-phosphoglycerate, from which the sugar is regenerated.  This cycle operates in plants, algae, cyanobacteria, some aerobic or facultative anaerobic Proteobacteria, CO-oxidizing mycobacteria and representatives of the genera Sulfobacillus (iron- and sulphur-oxidizing Firmicutes) and Oscillochloris (green sulphur bacteria).  An autotrophic symbiotic cyanobacterium conferred the CO2 fixation machinery on a eukaryotic cell giving rise to the chloroplasts of plant cells. The presence of the key enzyme, ribulose 1,5-bisphosphate carboxylase–oxygenase (RubisCO), is often considered to be synonymous with autotrophy.  Phylogenetic analysis and general considerations denote the Calvin cycle as a late innovation. Calvin–Benson–Bassham cycle Berg et al., 2010 6 Diagram illustrating the reactions of the Calvin–Benson–Bassham cycle. The corresponding enzymes are: 1: ribulose-1,5-bisphosphate carboxylase/oxygenase, 2: phosphoglycerate kinase, 3: glyceraldehyde-3-phosphate dehydrogenase, 4: triose phosphate isomerase, 5: fructose-bisphosphate aldolase, 6: fructose-1,6-bisphosphatase, 7: sedoheptulose bisphosphatase, 8: transketolase, 9: ribose-5-phosphate isomerase, 10: ribulose-5-phosphate 3-epimerase and 11: phosphoribulokinase. The directions of the arrows represent the direction of the pathway, and do not indicate that the enzyme reactions are irreversible. Calvin–Benson–Bassham cycle Sato & Atomi , 2010 7  In 1966, Arnon, Buchanan and co-workers proposed another autotrophic cycle for the green sulphur bacterium Chlorobium limicola, the reductive citric acid cycle (also known as the Arnon–Buchanan cycle).  This cycle is less energy-consuming than the Calvin cycle, involves enzymes that are sensitive to oxygen and is therefore found only in anaerobes or in aerobes growing at low oxygen tensions.  These include some Proteobacteria, green sulphur bacteria and microaerophilic bacteria of the early bacterial phylum Aquificae.  Initially, the reductive citric acid cycle was also proposed to operate in certain archaea (notably Thermoproteus neutrophilus), but recent findings refute this proposal. Berg et al., 2010 reductive tricarboxylic acid cycle 8 Diagram illustrating the reactions of the reductive tricarboxylic acid cycle. The corresponding enzymes are: 1: malate dehydrogenase, 2: fumarate hydratase, 3: fumarate reductase, 4: succinyl-CoA synthetase (acetyl-CoA: succinate CoA transferase is used in Desulfobacter hydrogenophilus), 5: oxoglutarate synthase, 6: isocitrate dehydrogenase, 7: aconitate hydratase, 8: ATP-citrate lyase, 9: pyruvate synthase, 10: pyruvate carboxylase, 11: phosphoenolpyruvate synthase and 12: phosphoenolpyruvate carboxylase. The directions of the arrows represent the direction of the pathway, and do not indicate that the enzyme reactions are irreversible. reductive tricarboxylic acid cycle Sato & Atomi , 2010 9  At the start of the 1980s, a third autotrophic pathway was found in certain Grampositive bacteria and methane-forming archaea, the reductive acetyl-coenzyme A (acetyl-CoA) or Wood–Ljungdahl pathway.  In these strict anaerobic organisms that now also include some Proteobacteria, Planctomycetes, spirochaetes and Euryarchaeota, one CO2 molecule is reduced to CO and one to a methyl group (bound to a carrier); subsequently, acetyl-CoA is synthesized from CO and the methyl group.  Although this pathway is the most energetically favourable autotrophic carbon fixation pathway, it is restricted to strictly anaerobic organisms.  Use of the acetyl-CoA pathway as an autotrophic, terminal electron-accepting process is the hallmark of acetogens. Other obligate anaerobic bacteriological groups, including methanogens and sulfate-reducing bacteria, also utilize the acetyl-CoA pathway for either catabolic or anabolic purposes. Berg et al., 2010; Drake et al., 1997 reductive acetyl CoA pathway 10 During the reductive synthesis of acetate via the acetyl-CoA pathway, six reducing equivalents are required for the fixation of CO2 to the methyl level, and two reducing equivalents are required for the fixation of CO2 to the carbonyl level. The overall conversion of substrate to product can be divided into oxidative and reductive processes. For example, the glycolytic utilization of glucose yields eight reducing equivalents per glucose consumed that are routed towards the reductive synthesis of acetate. reductive acetyl CoA pathway Drake et al., 1997 11 Drake et al., 2008 Homoacetogenic conversion of glucose to acetate. The two molecules of CO2 that are reduced to acetate in the acetyl-CoA pathway can be derived from exogenous CO2 rather than the CO2 that is produced via the decarboxylation of pyruvate (see text). ABBREVIATIONS: ATPSLP = ATP that is produced by substratelevel phosphorylation; [e−], reducing equivalent. reductive acetyl CoA pathway 12 reductive acetyl CoA pathway Schuchmann & Müller, 2014 13 14 Excursion: flavin-based electron bifuraction reductive acetyl CoA pathway Schuchmann & Müller, 2016 15 16 reductive acetyl CoA pathway Schuchmann & Müller, 2014 reductive acetyl CoA pathway Schuchmann & Müller, 2016 17 reductive acetyl CoA pathway 18 19 reductive acetyl CoA pathway Schuchmann & Müller, 2014 20 reductive acetyl CoA pathway Schuchmann & Müller, 2014 reductive acetyl CoA pathway Hypothetical scheme for the differential metabolic capacities of the acetogen Peptostreptococcus productus U-1. Abbreviations: LDH: lactate dehydrogenase; FBP: fructose- 1,6-bisphosphate; DHAP: dihydroxyacetone phosphate; GAP: glyceraldehyde-3-phosphate; OAA: oxaloacetate. Broken line indicates stimulation of lactacte dehydrogenase by fructose- 1,6-bisphosphate. Drake et al., 1997 21 reductive acetyl CoA pathway Drake et al., 1997, 2008 22 Representative oxidizeable substrates that can be used for the reductive synthesis of acetate via the acetyl-CoA ‘Wood/Ljungdahl’ pathway Drake et al., 1997 reductive acetyl CoA pathway 23 reductive acetyl CoA pathway Diagram illustrating the reactions of the reductive acetyl-CoA pathway. Coloured pathways indicate related enzymes involved in methanogenesis and acetogenesis. The corresponding enzymes are: 1: formate dehydrogenase, 2: formyl-H4FA-synthase, 3: methenyl-H4FA-cyclohydrolase, 4: methylene-H4FA-dehydrogenase, 5: methylene-H4FA-reductase, 6: methyltransferase, 7: carbon monooxide dehydrogenase/acetyl-CoA synthase, 8: phosphotransacetylase, 9: acetate kinase, 10: formylMF dehydrogenase, 11: formylMF:H4MPT formyltransferase, 12: N5, N10-methenyl-H4MPT cyclohydrolase, 13: N5, N10-methylene-H4MPT dehydrogenase, 14: N5, N10-methylene-H4MPT reductase, 15: N5-methyl-H4MPT:coenzyme M methyltransferase, 16: methyl-coenzyme M reductase and 17: heterodisulfide reductase. The directions of the arrows represent the direction of the pathway, and do not indicate that the enzyme reactions are irreversible. Abbreviations are: H4FA, tetrahydrofolate; MF, methanofuran; H4MPT, tetrahydromethanopterin; CoFe/S-P, corrinoid-iron sulfur protein. Sato & Atomi , 2010 24 Metabolic processes by which carbon monoxide is utilized by Methanosarcina acetivorans. (A) Oxidation of carbon monoxide; (B) reductive synthesis of acetate via the acetylCoA pathway; (C) production of methane. Parenthetical numbers identify enzymes that catalyze the indicated reactions: 1, CO dehydrogenase/acetyl-CoA synthetase; 2, formyl-methanofuran dehydrogenase; 3, formyl-methanofuran:H4MPT formyltransferase; 4, combined activities of methenyl-H4MPT cyclohydrolase, methylene-H4MPT dehydrogenase; methylene-H4MPT reductase; 5, phosphotransacetylase; 6, acetate kinase; 7, methyl-H4MPT:CoM methyltransferase; 8, combined activities of methyl-CoM reductase, membranous heterodisulfide reductase, and membranous F420H2 dehydrogenase complex (F420, coenzyme F420). ABBREVIATIONS: MF = methanofuran; H4MPT = tetrahydromethanopterin; HSCoM = coenzyme M. [The figure is based on information in Lessner et al. (further details on how these processes are coupled to the chemiosmotic conservation of energy can be found in this reference).] reductive acetyl-coenzyme A pathway 25 reductive acetyl-coenzyme A pathway Goyal et al., 2014 26 (1) Formylmethanofuran dehydrogenase (Fmd); (2) Formylmethanofuran:tetramethanopterin formyltransferase (Ftr), (3) methenyl-tetrahydromethanopterin cyclohydrolase (Mch), (4) methylene-tetrahydromethanopterin dehydrogenase (Hmd), (5) methylene-tetrahydromethanopterin reductase (Mer), (6) CO dehydrogenase-acetyl-CoA synthase Berg et al., 2010 reductive acetyl CoA pathway 27 reductive acetyl-coenzyme A pathway Metabolic pathway map of genes encoded in the SM1 euryarchaeal pangenome. Membrane proteins are indicated by symbols or protein structures of subunits (taken from Pfam database). Predicted proteins are indicated by numbers, which can be found in Supplementary Table 2. Grey numbers were found in SM1-MSI but not for SM1-CG. Specific enzymes were tested for presence of corresponding mRNA: Enzymes labeled in green were also detected in the mRNA pool of the biofilm, while enzymes with red color were absent in mRNA pool. Interrogation marks indicate potential metabolic pathway reactions that are likely present but are lacking evidence in the fragmented genomic bins. Probst et al., 2015 28  The 3-hydroxypropionate bicycle (3 hydroxypropionate/malyl CoA cycle) occurs in some green non-sulphur bacteria of the family Chloroflexaceae. This seems to be a singular invention, and the pathway has not been found elsewhere.  The conversion of acetyl-CoA plus two bicarbonates to succinyl-CoA uses the same intermediates as in the hydroxypropionate–hydroxybutyrate cycle, but most of the enzymes are completely different.  The regeneration of acetyl-CoA proceeds by the cleavage of malyl-CoA, yielding acetyl-CoA and glyoxylate. The assimilation of glyoxylate requires a second cycle (hence the name bicycle). 3-hydroxypropionate bicycle Berg et al., 2010 29 Diagram illustrating the reactions of the 3-hydroxypropionate bicycle (3-hydroxypropionate/malyl-CoA cycle). The corresponding enzymes are: 1: acetyl-CoA carboxylase, 2: malonyl-CoA reductase, 3: propionyl-CoA synthase, 4: propionyl-CoA carboxylase, 5: methylmalonyl-CoA epimerase, 6: methylmalonyl-CoA mutase, 7: succinyl-CoA:(S)-malate-CoA transferase, 8: succinate dehydrogenase, 9: fumarate hydratase, 10-1, 10-2, 10-3: (S)-malyl-CoA/β-methylmalyl-CoA/(S)-citramalyl-CoA (MMC) lyase, 11: mesaconyl-C1-CoA hydratase (β-methylmalyl-CoA dehydratase), 12: mesaconyl-CoA C1-C4 CoA transferase and 13: mesaconyl-C4-CoA hydratase. The directions of the arrows represent the direction of the pathway, and do not indicate that the enzyme reactions are irreversible. 3-hydroxypropionate bicycle Sato & Atomi , 2010 30 Pathways of autotrophic CO2 fixation in crenarchaeota. The dicarboxylate–hydroxybutyrate cycle functions in Desulfurococcales and Thermoproteales (a) and the hydroxypropionate–hydroxybutyrate cycle functions in Sulfolobales and Thaumarchaeota (b). Note that succinyl-coenzyme A (succinyl-CoA) reductase in Thermoproteales and Sulfolobales uses NADPH and reduced methyl viologen (possibly as a substitute for reduced ferredoxin) in Desulfurococcales. In Sulfolobales, pyruvate might be derived from succinyl-CoA by C4 decarboxylation. CoASH, coenzyme A; Fdred 2–, reduced ferredoxin; Fdox, oxidized ferredoxin; PEP, phosphoenolpyruvate. Berg et al., 2010 pathways of autotrophic CO2 fixation in crenarchaeota dicarboxylate–hydroxybutyrate cycle hydroxypropionate–hydroxybutyrate cycle 31  The 3 hydroxypropionate/4 hydroxybutyrate cycle (hydroxypropionate– hydroxybutyrate cycle) was discovered by Berg et al. in 2007.  The hydroxypropionate–hydroxybutyrate cycle occurs in Sulfolobales and Thaumarchaeota. Although some of the intermediates and the carboxylation reactions are the same as in the 3-hydroxypropionate bicycle in Chloroflexaceae, the archaeal cycle probably has evolved independently. Berg et al., 2007, 2010 3 hydroxypropionate/4 hydroxybutyrate cycle 32 3 hydroxypropionate/4 hydroxybutyrate cycle Proposed autotrophic 3-hydroxypropionate/4hydroxybutyrate cycle in M. sedula. Reactions of the cycle are shown. Enzymes: 1, acetyl-CoA carboxylase; 2, malonyl-CoA reductase (NADPH); 3, malonate semialdehyde reductase (NADPH); 4, 3-hydroxypropionyl-CoA synthetase (AMP- forming); 5, 3-hydroxypropionyl-CoA dehydratase; 6, acryloyl-CoA reductase (NADPH); 7, propionyl-CoA carboxylase; 8, methylmalonyl-CoA epimerase; 9, methylmalonyl-CoA mutase; 10, succinyl-CoA reductase (NADPH); 11, succinate semialdehyde reductase (NADPH); 12, 4-hydroxybutyryl-CoA synthetase (AMP- forming); 13, 4-hydroxybutyryl-CoA dehydratase; 14, crotonyl-CoA hydratase; 15, 3-hydroxybutyryl-CoA dehydrogenase (NAD+); 16, acetoacetyl-CoA β-ketothiolase. The proposed pathway of glyceraldehyde 3phosphate synthesis from acetyl-CoA and CO2 is also shown. Enzymes: 17, pyruvate synthase; 18, pyruvate, water dikinase [phosphoenolpyruvate (PEP) synthase]; 19, enolase; 20, phosphoglycerate mutase; 21, 3-phosphoglycerate kinase; 22, glyceraldehyde 3-phosphate dehydrogenase. Berg et al., 2007 33 Reactions of the crenarchaeal and thaumarchaeal variants of the HP/ HB cycle. The reactions determining energy efficiency of the crenarchaeal (M. sedula) cycle are shown in green, and those for the N. maritimus variant are shown in red. Reactions common to both are shown in black. Note that although the two pathways have similar reactions and intermediates, they are significantly different in energy efficiency and evolved independently in Crenarchaeota and Thaumarchaeota. The numbers in square brackets represent moles of high-energy anhydride bonds of ATP required to form 1 mol of the corresponding central precursor metabolites. Enzymes as numbered in circles are: 1, 3-hydroxypropionyl-CoA synthetase (ADP-forming); 2, 3-hydroxypropionyl-CoA synthetase (AMP-forming); 3, 3-hydroxypropionyl-CoA dehydratase; 4, 4-hydroxybutyryl-CoA synthetase (ADP-forming); 5, 4-hydroxybutyryl-CoA synthetase (AMP-forming); 6, 4-hydroxybutyryl-CoA dehydratase; 7, crotonyl-CoA hydratase; 8, succinyl-CoA synthetase (ADP-forming); 9, succinic semialdehyde dehydrogenase; 10, pyruvate-phosphate dikinase; and 11, pyruvatewater dikinase. Catalytic properties of recombinant enzymes Könnecke et al,. 2015 3 hydroxypropionate/4 hydroxybutyrate cycle 34 3 hydroxypropionate/4 hydroxybutyrate cycle 35 3 hydroxypropionate/4 hydroxybutyrate cycle 36 0,0 1000,0 2000,0 3000,0 4000,0 5000,0 6000,0 0,0 50,0 100,0 150,0 200,0 250,0 c(NO2 -)[µmolL-1] time [h] c(NO2-) Results of quadruplicate batch fermentation of N. viennensis in a bioreactor system with continuous supply of 20% CO2, 21% O2, rest N2. Fed-batch fermentation of N. viennensis 37  The dicarboxylate/4 hydroxybutyrate cycle was discovered by Huber et al. in 2008.  The dicarboxylate–hydroxybutyrate cycle occurs in the anaerobic crenarchaeal orders Thermoproteales and Desulfurococcales. Huber et al., 2008 dicarboxylate/4 hydroxybutyrate cycle 38 Diagram illustrating the reactions of the dicarboxylate/4-hydroxybutyrate cycle. The corresponding enzymes are: 1: pyruvate synthase, 2: pyruvate:water dikinase, 3: phosphoenolpyruvate carboxylase, 4: malate dehydrogenase, 5: fumarate hydratase, 6: fumarate reductase, 7: succinate thiokinase, 8: succinyl-CoA reductase, 9: succinate semialdehyde reductase, 10: 4-hydroxybutyryl-CoA synthetase, 11: 4-hydroxybutyryl-CoA dehydratase, 12: crotonyl-CoA hydratase, 13: 3-hydroxybutyryl-CoA dehydrogenase and 14: acetoacetyl-CoA β-ketothiolase. The directions of the arrows represent the direction of the pathway, and do not indicate that the enzyme reactions are irreversible dicarboxylate/4 hydroxybutyrate cycle Sato & Atomi, 2010 39 Kono et al., 2016 Reductive hexulose-phosphate pathway Kono et al., 2016 Reductive hexulose-phosphate pathway Structural comparison of archaeal and photosynthetic PRKs. (a,b) Topology diagrams of the folding patterns in protomers of (a) M. hungatei PRK (MhPRK) and (b) R. sphaeroides PRK (RsPRK) are shown in yellow and blue, respectively. a-helices are denoted by cylinders, b-sheets by arrows and connecting loops by lines. Positions in the sequence that start and end each major secondary structural element are shown. (c,d) Ribbon diagrams of (c) MhPRK (Protein Data Bank (PDB) ID 5B3F) monomer and (d) RsPRK (PDB ID 1A7J) monomer. Sulphate ions are bound to the active site of MhPRK and RsPRK. Disordered regions of missing electron density are shown as dots for residues for 156–163 in MhPRK, residues 15–17 corresponding to a part of the P-loop, and residues 100–105 in RsPRK. (e,f) Active-site structures of (e) MhPRK and (f) RsPRK. Side chains of residues involved in ATPand Ru5P-binding are shown as pink and green sticks, respectively, with oxygen atoms in red and nitrogen atoms in blue. Sulphate ions are shown in MhPRK and RsPRK, with sulphur atoms in green, and oxygen and nitrogen atoms in the same colours as those in residue side chains. Small red balls represent water molecules. Disordered regions are shown as dots, and P-loop containing residues involved in ATP binding are shown in pink. Dotted lines show interactions of active site sulphate ions. Kono et al., 2016 Reductive hexulose-phosphate pathway Kono et al., 2016 Reductive hexulose-phosphate pathway Kono et al., 2016 Reductive hexulose-phosphate pathway Bar-Even et al., 2012 CO2 fixation pathways 45 Oxygen, metals and supply of C1 compounds Autotrophic Euryarchaeota are strictly confined to anoxic conditions, generally specialized in metabolizing C1 compounds and/or acetate, and their energy metabolism has low ATP yields. Therefore, they need much of the C1-transforming machinery for their energy metabolism. The reductive acetyl-coenzyme A (acetyl-CoA) pathway ideally copes with such constraints. Also, essential metals are more available under anoxic conditions owing to the higher solubility of the reduced forms of most metals. In Crenarchaeota, the oxygen-sensitive dicarboxylate–hydroxybutyrate cycle is restricted to the anaerobic Thermoproteales and Desulfurococcales, whereas the oxygen-insensitive hydroxypropionate– hydroxybutyrate cycle is restricted to the mostly aerobic Sulfolobales and Thaumarchaeota. The two lifestyles presuppose different electron donors with different redox potentials and different oxygen sensitivity of cofactors and enzymes. In a nutshell, energy cost-effective but oxygen-sensitive mechanisms cannot exist in aerobes because the enzymes would be inactivated by oxygen; and not all anaerobes have C1 substrates at their disposal. Berg et al., 2010 Benefits of the different autotrophic pathways 46 Energy demands The different pathways require different amounts of ATP to make the cellular precursor metabolites. The costs for synthesizing all auxiliary, CO2 fixation-related enzymes also differ, which might determine the energy costs involved. The synthesis of the catalysts itself can require a huge amount of energy as well as nitrogen and sulphur sources, especially if the pathways involve many auxiliary enzymes. Carboxylases with low catalytic efficiency must be synthesized in large amounts, as is the case for ribulose 1,5-bisphosphate carboxylase– oxygenase (RUBISCO). So, energy limitation exerts a strong selective pressure in favour of energy-saving mechanisms, and the energy costs are largely spent for the synthesis of autotrophy-related enzymes. Berg et al., 2010 Benefits of the different autotrophic pathways Könnicke et al., 2015 47 Properties of the CO2 fixation pathways Berg et al., 2010; Bar-Even et al., 2012 48 Metabolic fluxes In bacteria and archaea, the need for sugar phosphates in the biosynthesis of cell walls is lower than in plants, which also synthesize huge amounts of cellulose and lignin that is derived from erythrose 4-phosphate and phosphoenolpyruvate (PEP). The main metabolic fluxes are diverted from acetyl-CoA, pyruvate, oxaloacetate and 2-oxoglutarate, and their synthesis from 3phosphoglycerate is partly connected with a loss of CO2. Therefore, in bacteria autotrophic pathways directly yielding acetyl-CoA are more economical. Still, most facultative aerobic bacteria use the Calvin cycle, the regulation of which is almost detached from the central carbon metabolism and therefore may be particularly robust. Berg et al., 2010 Benefits of the different autotrophic pathways 49 CO2 species As the bicarbonate (HCO3 –) concentration in slightly alkaline water is much higher than the concentration of dissolved CO2, autotrophs might profit from using bicarbonate instead of CO2. The usage of bicarbonate is a special feature of PEP carboxylase and biotin-dependent carboxylases (that is, acetyl-CoA–propionyl-CoA carboxylase). This property of PEP carboxylase is used in plants in the crassulacean acid and C4 metabolism to increase the efficiency of photosynthesis. The same might be true for acetyl-CoA–propionylCoA carboxylase, and the higher bicarbonate concentration could potentially make up for a lower bicarbonate affinity. Berg et al., 2010 Benefits of the different autotrophic pathways 50 Co-assimilation of organic compounds Many autotrophic bacteria and archaea living in aquatic habitats probably encounter carbon oligotrophic conditions and grow as mixotrophs. Co-assimilation of traces of organic compounds might pay off. A complete or even a rudimentary hydroxypropionate–hydroxybutyrate cycle, for instance, allows the co-assimilation of numerous compounds. These include fermentation products and 3-hydroxypropionate, an intermediate in the metabolism of the ubiquitous osmoprotectant dimethylsulphoniopropionate. It is possible that various widespread marine aerobic phototrophic bacteria have genes encoding a rudimentary 3-hydroxypropionate cycle for that purpose. Similarly, the dicarboxylate– hydroxybutyrate cycle allows the co-assimilation of dicarboxylic acids and substrates that are metabolized through acetyl-CoA. Berg et al., 2010 Benefits of the different autotrophic pathways 51 Microbial formate utilization Crable et al. , 2011 52 Bar-Even, 2016 Microbial formate utilization 53 Microbial formate utilization Reaction conditions: pH 7,0; 1M concentration, gases at 1 atm Schink et al., 2017 54 Microbial formate utilization Schink et al., 2017 55 Microbial formate utilization Schink et al., 2017 56 Bar-Even et al., 2012 Microbial formate utilization 57 Microbial formate utilization Bar-Even, 2016 Reductive acetyl-CoA pathway 58 Microbial formate utilization Bar-Even, 2016 Serine cycle 59 Microbial formate utilization Bar-Even, 2016 Reductive glycine pathway 60 61 Kottenhahn et al., 2018 H2 production from formate Microbial formate utilization 62 Ceccaldi et al., 2017 A. woodii 63 Kottenhahn et al., 2018 H2 production from formate 64 Kottenhahn et al., 2018 H2 production from formate Rittmann et al. , 2015 Microbial formate utilization 65 66 Rittmann et al. , 2015 Microbial formate utilization 67 Kim et al., 2010 H2 production from formate 68 Rittmann et al. (2012) Microbial Cell Factories Rittmann et al. (2015) Biotechnology Advances Ergal et al. (2018) Biotechnology Advances HER: hydrogen evolution rate H2 production from formate Costa et al., 2010 Microbial formate utilization 69 Microbial formate utilization Costa et al., 2013(c) 70 Microbial formate utilization Costa et al., 2013(a,b) 71 Microbial formate utilization Lupa et al., 2008 72 Microbial formate utilization Lupa et al. , 2008 73 Microbial formate utilization Crable et al. , 2011 74 C-cycle Offre et al., 2014 Red arrows: metabolic stepts related to archaea and bacteria; Organge arrows: exclusive metabolic routes only related to arcahea Modes of microbial anaerobic methane oxidation. There are four known ways in which microorganisms achieve anaerobic oxidation of methane (AOM). Two of these (a, b) are thought to rely on obligate associations between two or more microbial partners, one of which performs oxidation and the other reduction; in the other two cases (c, d), a single microorganism performs both reactions. a, Anaerobic methanotrophic archaea (ANMEs) oxidize methane (CH4) and convert it to carbon dioxide and water, in cooperation with sulphate-reducing bacteria, which convert sulphate (SO42−) to hydrogen sulphide (H2S). The mechanism of energy exchange between the ANMEs and the sulphate-reducing bacteria is unknown. b, The oxidation of methane to CO2 by ANMEs is coupled to the reduction of metal oxides, whereby metals such as manganese (Mn) or iron (Fe) are reduced to the +2 oxidation state. c, The bacterium Methoxymirabilis oxyfera converts nitrite (NO2−) to nitric oxide (NO) and then dismutates (splits) NO into nitrogen and oxygen as diatomic gases. The bacterium then uses the resulting O2 to support methane oxidation. d, Milucka et al. show that some ANMEs oxidize methane (as in a) but also reduce sulphate to zero-valent sulphur (S0), which they produce in the form of disulphide (HS2−). The disulphide can be used by associated bacteria, Deltaproteobacteria, to yield sulphide (HS−) and sulphate, but this is an association of convenience, rather than necessity. Joye, 2012 Microbial methane oxidation Microbial methane oxidation Milucka et al., 2008 Disulphide disproportionation by Isis enrichment culture. a, Linear correlation between sulphate and sulphide production in the Isis enrichment culture incubated with 13CO2 and colloidal sulphur in the absence of methane over the course of 70 days (n=51) follows the 7:1 ratio inferred for disulphide disproportionation. b, Overlay of nanoSIMS12C14Nimage (red) and fluorescence image of microorganisms from the enrichment culture stained with the DSS-658 probe (green). c, The 13C/12C image shows that the DSS cells (several of which are shown by white outlines) are enriched in 13C. Coloured scale below panel c shows ratio of the respective ions. d, Production of sulphide and sulphate (based on 35S accumulation in sulphate after the addition of 35Ssulphide) under AOM conditions (n=51; this experiment was repeated twice). Microbial methane oxidation Milucka et al., 2008 Microbial methane oxidation Revised model of anaerobic oxidation of methane coupled to sulphate reduction. ANME-2 oxidize methane with a concomitant reduction of sulphate to zero-valent sulphur (S0, elemental sulphur) that is partially deposited or bound intracellularly. Produced S0 is exported or diffuses outside the cell where it reacts with sulphide to form polysulphides (disulphide, among others). Disulphide is taken up by the associated Deltaproteobacteria and is disproportionated to sulphate and sulphide. Sulphate produced during disproportionation might be re-used by the ANME and the ANME may also reduce some of the sulphate all the way to sulphide (grey dotted line). Dark circles in the bacteria represent intracellular precipitates rich in iron and phosphorus. Milucka et al., 2008 Microbial methane oxidation Significant pathways of Methylomirabilis oxyfera. Canonical pathways of denitrification (a), aerobic methane oxidation (b) and proposed pathway of methane oxidation with nitrite (c). narGHJI, nitrate reductase; napABCDE, periplasmic nitrate reductase; nirSJFD/GH/L, nitrite reductase; norZ, nitric oxide reductase; nosDFYLZ, nitrous oxide reductase; pmoCAB, particulate methane monooxygenase; mxaFJGIRSACKL/DE, methanol dehydrogenase; fae, formaldehydeactivating enzyme; mtdB, methylenetetrahydromethanopterin (H4MPT) dehydrogenase; mch, methenylH4MPT cyclohydrolase; fhcABCD, formyltransferase/hydrolase; fdhABC, formate dehydrogenase. Genes in red are absent from the genome, those in blue are present in the genome and those genes in green are present in both the proteome and the genome. Asterisk, H4MPT-dependent reactions involve the intermediates methylene H4MPT, methenyl-H4MPT and formyl- H4MPT. Ettwig et al. , 2010 Microbial methane oxidation Ettwig et al. , 2010 Thauer and Shima, 2006 Haroon et al. , 2013 Microbial methane oxidation Haroon et al. , 2013 Key carbon and nitrogen transformations in ‘Methanoperedens nitroreducens’. Reverse methanogenesis pathway in Ca. ‘M. nitroreducens’ coupled by an unknown electron carrier to nitrate reduction. Highly expressed genes are shown in red, indicating that the complete reverse methanogenesis pathway and nitrate reduction genes were active in the bioreactor. Increasing line thickness indicates increasing absolute gene expression values. FPKG (fragments mapped per kilobase of gene length) is a measure of normalized gene expression. Microbial methane oxidation Haroon et al. , 2013 Microbial methane oxidation Haroon et al. , 2013 Microbial methane oxidation