<<

Delft University of Technology

Discovery of extremely halophilic, methyl-reducing euryarchaea provides insights into the evolutionary origin of

Sorokin, DImitry Y.; Makarova, Kira S.; Abbas, Ben; Ferrer, Manuel; Golyshin, Peter N.; Galinski, Erwin A.; Ciordia, Sergio; Mena, María Carmen; Merkel, Alexander Y.; Wolf, Yuri I. DOI 10.1038/nmicrobiol.2017.81 Publication date 2017 Document Version Accepted author manuscript Published in Nature Reviews. Microbiology

Citation (APA) Sorokin, DI. Y., Makarova, K. S., Abbas, B., Ferrer, M., Golyshin, P. N., Galinski, E. A., Ciordia, S., Mena, M. C., Merkel, A. Y., Wolf, Y. I., Van Loosdrecht, M. C. M., & Koonin, E. V. (2017). Discovery of extremely halophilic, methyl-reducing euryarchaea provides insights into the evolutionary origin of methanogenesis. Nature Reviews. Microbiology, 2(8), [17081]. https://doi.org/10.1038/nmicrobiol.2017.81 Important note To cite this publication, please use the final published version (if applicable). Please check the document version above.

Copyright Other than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons. Takedown policy Please contact us and provide details if you believe this document breaches copyrights. We will remove access to the work immediately and investigate your claim.

This work is downloaded from Delft University of Technology. For technical reasons the number of authors shown on this cover page is limited to a maximum of 10. 1 Discovery of extreme halo(alkali)philic, thermophilic, methyl-reducing 2 methanogenic euryarchaea 3 4 Dimitry Y. Sorokin1,2*, Kira S. Makarova3, Ben Abbas2, Manuel Ferrer4, Peter N. Golyshin5, 5 Erwin A. Galinski6, Sergio Ciordia7, María Carmen Mena7, Alexander Y. Merkel1, Yuri I. 6 Wolf3, Mark C.M. van Loosdrecht2, Eugene V. Koonin3* 7 8 1Winogradsky Institute of Microbiology, Centre for Biotechnology, Russian Academy of Sciences, Moscow, 9 Russia; 10 2Department of Biotechnology, Delft University of Technology, Delft, The Netherlands; 11 3National Center for Biotechnology Information, National Library of Medicine, National Institutes of Health, 12 Bethesda, MD, USA; 13 4Institute of Catalysis, CSIC, Madrid, Spain; 14 5School of Biological Sciences, Bangor University, Gwynedd, UK 15 6Institute of Microbiology and Biotechnology, Rheinische Friedrich-Wilhelms University, Bonn, Germany 16 7Proteomics Facility, Centro Nacional de Biotecnología, CSIC, Madrid, Spain 17 18 19 *Corresponding authors: 20 Dimitry Y. Sorokin: [email protected]; [email protected] 21 Eugene V. Koonin: [email protected] 22

1

23 Methanogenic are major players in the global carbon cycle and in the biotechnology of 24 anaerobic digestion. The phylum includes diverse groups of that are 25 interspersed with non-methanogenic lineages. So far methanogens inhabiting hypersaline 26 environments have been identified only within the order . We report the 27 discovery of a deep phylogenetic lineage of extremophilic methanogens in hypersaline lakes, and 28 present analysis of two nearly complete from this group. Within the phylum 29 Euryarchaeota, these isolates form a separate, class-level lineage "Methanonatronarchaeia" that 30 is most closely related to the class Halobacteria. Similar to the Halobacteria, 31 "Methanonatronarchaeia" are extremely halophilic and do not accumulate organic 32 osmoprotectants. These methanogens are heterotrophic methyl-reducers that utilize C1- 33 methylated compounds as electron acceptors and formate or as electron donors. The 34 genomes contain an incomplete and apparently inactivated set of genes encoding the upper 35 branch of methyl group oxidation to CO2 and membrane-bound heterosulfide reductase and 36 cytochromes. These features differentiates "Methanonatronarchaeia" from all known methyl- 37 reducing methanogens. The high intracellular concentration of potassium implies that 38 "Methanonatronarchaeia" employ the "salt-in" osmoprotection strategy. The discovery of 39 extremely halophilic, methyl-reducing methanogens related to sheds new light on 40 the origin of methanogenesis and shows that the strategies employed by methanogens to thrive in 41 salt-saturating conditions are not limited to the classical methylotrophic pathway.

2

42 Introduction 43 44 Methanogenesis is one of the key terminal anaerobic processes of the biogeochemical carbon

45 cycle both in natural ecosystems and in industrial biogas production plants 1,2. Biomethane is a major

46 contributor to global warming 3. Methanogens comprise four classes, "",

47 , and , and part of the class , within the

48 archaeal phylum Euryarchaeota4-7. The recent metagenomic discovery of putative methyl-reducing

49 methanogens in the Candidate phyla "Bathyarchaeota" 8 and "Verstaraetearchaeota" 9 indicates that

50 methanogenesis might not be limited to Euryarchaeota.

1,2 51 Three major pathways of methanogenesis are known : hydrogenotrophic (H2, formate and

52 CO2/bicarbonate as electron acceptor), methylotrophic (dismutation of C1 methylated compounds to

53 and CO2) and acetoclastic (dismutation of into methane and CO2). In the

54 hydrogenotrophic pathway, methane is produced by sequential 6-step reduction of CO2. In the

55 methylotrophic pathway, methylated C1 compounds, including , and

56 methylsulfides, are first activated by specific methyltransferases. Next, one out of four methyl groups

57 is oxidized through the same reactions as in the hydrogenotrophic pathway occurring in reverse, and

58 the remaining three groups are reduced to methane. In the acetoclastic pathway, methane is produced

59 from the methyl group after activation of acetate. The only enzyme that is uniquely present in all three

60 types of methanogens is methyl-CoM reductase, a Ni-corrinoid protein catalyzing the last step of

61 methyl group reduction to methane 10-12.

62 The recent discovery of methanogens among Thermoplasmata 5,13-15 drew attention to the

63 fourth, methyl-reducing, pathway, previously characterized in Methanosphaera stadtmanae

16-20 64 ( Methanobacteria) and Methanomicrococcus blatticola ("Methanomicrobia") . In this pathway, C1

65 methylated compounds are used only as electron acceptors, whereas H2 serves as electron donor. In the

66 few known representatives, the genes for methyl group oxidation to CO2 are either present but inactive

67 ( Methanosphaera) 16 or completely lost (Thermoplasmata methanogens) 6-7. Recent metagenomic

68 studies have uncovered three additional, deep lineages of potential methyl-reducing methanogens,

69 namely, Candidate class "Methanofastidiosa" within Euryarchaeota 21 and Candidate phyla

3

70 "Bathyarchaeota" and "Verstraetearchaeota" 8,9, supporting the earlier hypothesis that this is an

71 independently evolved, ancient pathway 22.

72 The classical methylotrophic pathway of methanogenesis that has been characterized in

73 moderately halophilic members of Methanosarcinales 23, apparently dominates in hypersaline

74 conditions 23-25. In contrast to the extremely halophilic haloarchaea, these microbes only tolerate

75 saturated salt conditions but optimally grow at moderate salinity (below 2-3 M Na+) using organic

76 compounds for osmotic balance ("salt-out" strategy) 26,27.

77 Our recent study of methanogenesis in hypersaline soda lakes identified methylotrophic

78 methanogenesis as the most active pathway. In addition, culture-independent analysis of the mcrA

79 gene, a unique marker of methanogens, identified a deep lineage that is only distantly related to other

80 methanogens 28. We observed no growth of these organisms upon addition of substrates for the

81 classical methanogenic pathways and concluded that they required distinct growth conditions. Here we

82 identify such conditions and describe the discovery and physiological, genomic and phylogenetic

83 features of a previously overlooked group of extremely halophilic, methyl-reducing methanogens.

84

85 Discovery of an unknown deep lineage of extremely halophilic methanogens in hypersaline lakes

86 Sediment stimulation experiments

87 Two deep-branching mcrA sequences have been previously detected in sediments from hypersaline

88 soda lakes in south-eastern Siberia 28. Attempts to stimulate the activity of these uncharacterized,

89 dormant methanogens by variation of conditions (temperature, pH and salinity) and substrates elicited

90 a positive response at extreme salinity (4 M Na+), pH (9.5-10), elevated temperature (above 48-55oC)

91 and in the presence of methylotrophic substrates together with formate or H2 (the combination used in

92 the methyl-reducing pathway). The typical response involved a pronounced increase in methane

93 production upon combining methyl compounds with formate or H2 (less active) compared to single

94 substrates (Supplementary Figure 1 a). The mcrA profiling of such incubations revealed two distinct

95 clusters closely related to the previously detected deep methanogenic lineage 28 (Supplementary

96 Figure 2).

4

97 The same approach was used with sediment slurries from hypersaline lakes with neutral pH

98 (with no previous evidence of the presence of methyl-reducing methanogens). In this case, enhanced

99 methane production under methyl-reducing conditions (MeOH/trimethylamine + formate) was also

100 observed at elevated temperatures (Supplementary Figure 1 b,c). The mcrA profiles indicated that

101 typical halophilic methylotrophic methanogens ( and ) were

102 outcompeted at high temperature (50-60oC) by unknown, extremely halophilic methyl-reducers which

103 formed a sister clade to the sequences from methyl-reducing incubations of soda lakes sediments in

104 the mcrA tree (Supplementary Figure 2).

105

106 Cultivation of the extremely halophilic methyl-reducing methanogens

107 The active sediment incubations from hypersaline lakes (Supplementary Table 1) were used as an

108 enriched source to obtain the methyl-reducing methanogens in laboratory culture using synthetic

109 media with 2-4 M Na+, pH 7 (for salt lakes) or 9.5-10 (for soda lakes), supplemented with

110 MeOH/formate or trimethylamine (TMA)/formate and incubated at 48-60oC. Methane formation was

111 observed only at extreme salinity, close to saturation (4 M total Na+), but ceased after the original

112 sediment inoculum was diluted by 2-3 consecutive 1:100 transfers. Addition of colloidal FeSxnH2O

113 (soda lakes) or sterilized sediments (salt lakes), combined with filtration through 0.45 μm filters and

114 antibiotic treatment, yielded a pure culture from Siberian soda lakes (strain AMET1 [Alkaliphilic

115 Methylotrophic Thermophilic]), and 10 additional pure AMET cultures from hypersaline alkaline

116 lakes in various geographic locations. A similar approach resulted in three highly enriched cultures at

117 neutral pH from salt lakes (HMET [Halophilic Methylotrophic Thermophilic] cultures)

118 (Supplementary Table 2). Phylogenetic analysis of the marker genes showed that AMET and HMET

119 formed two potential -level groups that shared 90% 16S rRNA gene sequence identity.

120 121 Microbiological characteristics of the methyl-reducing methanogens 122 123 morphology and composition

124 Both AMET and HMET possess small coccoid cells that are motile, in the case of AMET, and lack

125 F420 autofluorescence that is typical of most methanogens. A thin, single-layer was present in

5

126 both groups (Figure 1; Supplementary Figure 3). At salt concentration below 1.5 M total Na+, the

127 cells lost integrity.

128 The extreme halophily of the discovered methanogens is unprecedented. The salt-tolerant

129 isolated so far from hypersaline habitats, such as Methanohalobium,

130 Methanohalophilus and , all accumulate organic osmolytes ("salt-out"

131 osmoprotection). In contrast, no recognizable organic osmolytes were detected in AMET1 cells that,

132 instead, accumulated high intracellular concentrations of potassium [5.5 μmol/g protein or 2.2 M,

133 assuming the cell density of 1.2 mg/ml for haloarchaea 29 and the measured protein content of 30%].

134 This concentration is twofold lower than that normally observed inside the cells of haloarchaea (12- 13

135 μmol/g protein) but close to that of Halanaerobium (6.3 μmol/g protein), both of which have been

136 shown to employ the "salt-in" osmoprotection strategy 30,31. Furthermore, half of the sodium in the

137 medium is present in the form of carbonates, which possess exactly twofold less osmotic activity than

138 NaCl, resulting in decreased total osmotic pressure, and accordingly, a lower intracellular

139 concentration of osmolytes in extreme natronophiles 32. This finding suggests that the extremely

140 halophilic methyl-reducing methanogens rely on potassium as the major osmolyte.

141 The AMET cell pellets were pinkish in color, suggestive of the presence of cytochromes

142 which was confirmed by difference spectra of a cell-free extract from AMET1 that showed peaks

143 characteristic of b-type cytochromes (Supplementary Figure 4 a). Given that the cytochrome-

144 containing methanogens of the order Methanosarcinales also synthesize the electron-transferring

145 quinone analogue methanophenazine 33, we attempted to detect this compound in AMET1. Indeed, two

146 yellow-colored autofluorescent hydrophobic fractions were recovered from the AMET1 cells, with

147 main masses of 562 and 580 Da, which behaved similar to methanophenazine from

148 (mass 532 Da) upon chemical ionization (sequential cleavage of the 68 Da mass isoprene unit)

149 (Supplementary Figure 4 b).

150

151 Growth physiology

152 Both AMET and HMET are methyl-reducing heterotrophic methanogens utilizing C1-methyl

153 compounds as e-acceptor, formate or H2 as e-donor, and yeast extract or acetate as the C-source. 6

154 Growth of both groups of organisms was stimulated by addition of external CoM (up to 0.1 mM).

155 Despite the general metabolic similarity, the AMET cultures grew and survived long storage much

156 better than the HMET cultures. The AMET cultures grew best with MeOH as acceptor and formate as

157 donor (Figure 2 a). Apart from MeOH, slower growth was also observed with methylamines and

158 dimethylsulfide (Figure 2 b). In sharp contrast to the known methyl-reducing methanogens, H2 was

159 less effective as the electron donor.

160 Both groups grew optimally around 50oC, with the upper limit at 60oC (Figure 2 c,

161 Supplementary Figure 5). The AMET isolates were obligate alkaliphiles, with optimum growth at

162 pH 9.5-9.8 (Figure 2 d), whereas the HMET cultures had an optimum at pH 6.8-7. The organisms of

163 both groups showed the fastest growth and the highest activity at salt-saturating conditions, and thus

164 qualified as extreme halo(natrono)philes (Figure 2 e,f).

165

166 Effect of iron sulfides on growth and activity of AMET1

167 Apart from hydrotroilite (FeSxnH2O), AMET1 also grew, albeit less actively, in the presence of

168 cristalline FeS, and yet less actively, with pyrite (FeS2). No other forms of reduced iron minerals tested

169 (olivine, FeCO3, magnetite, ferrotine (FeSn) or various iron(II) silicates could replace FeS.

170 Furthermore, methanogenic activity of resting cells depleted for FeS showed dependence on FeS

171 addition (Figure 3). No methane was formed in the absence of either methyl acceptors or formate/H2,

172 suggesting that Fe2+ likely served as a catalyst or regulator rather than a direct e-donor. The specific

173 cause(s) of the dependence of AMET growth on iron (II) sulfides remains to be identified.

174

175 Comparative genomic analysis

176 General characteristics

177 The general genome characteristics of AMET1 and HMET1 are given in Table 1. Based on

178 analysis of 218 core arCOGs 34, both genomes are nearly complete, with two genes missing from this

179 list in AMET1 and three in HMET1. Two of these genes are missing in both genomes (prefoldin

180 paralog GIM5 and deoxyhypusine synthase DYS1), suggesting that they were lost in the common

181 ancestor (Supplementary Table 3). The presence of tRNAs for all amino acids is another indication

7

182 of genome completeness. The high coverage of the AMET1 and HMET1 genomes by arCOGs implies

183 that the unique phenotype of these organisms is supported largely by the already well-sampled part of

184 the archaeal gene pool.

185

186 Phylogenetic analysis and

187 A concatenated alignment of the 56 ribosomal proteins that are universally conserved in

188 complete archaeal genomes 35 including AMET1 and HMET1 was used for maximum likelihood tree

189 reconstruction (Figure 4 a, Supplementary Table 3, Supplementary Data 1). Both AMET1 and

190 HMET1 belong to a distinct clade, a sister taxon to the class Halobacteria, with 100% bootstrap

191 support (Figure 4 a). The 16S rRNA gene tree suggests that both organisms belong to the uncultured

192 SA1 group that was first identified in the brine-seawater interface of the Shaban Deep in the Red Sea36

193 and subsequently in other hypersaline habitats37 (Supplementary Figure 6). According to the rRNA

194 phylogeny, the group that includes AMET1 and HMET1 is well separated from the other classes in the

195 phylum Euryarchaeota, both methanogenic and non-methanogenic. The 16S rRNA sequences of

196 these organisms are equally distant from all classes in Euryarchaeota and fall within the range

197 of recently recommended values (80-86%) for the class level classification 38. Together, these

198 findings appear to justify classification of the SA1 group, including the AMET and HMET lineages, as

199 a separate euryarchaeal class "Methanonatronarchaeia". This class would be represented by two

200 distinct genera and "Methanonatronarchaeum thermophilum" (AMET) and 'Candidatus

201 Methanohalarchaeum thermophilum' (HMET).

202

203 Comparative genomic analysis and reconstruction of main evolutionary events

204 Using arCOG assignments and the results of previous phylogenomic analysis 39, we

205 reconstructed the major evolutionary events in the history of AMET1, HMET1 and Halobacteria

206 (Figure 4 a and Supplementary Table 4). This reconstruction indicates that evolution of the

207 HMET1-AMET1 lineage was dominated by gene loss, whereas Halobacteria acquired most of their

208 gene complement after the divergence from "Methanonatronarchaeia". As shown previously, the

8

209 common ancestor of Methanomicrobia and Halobacteria was a 39. The key genes coding

210 for components of the protein complexes involved in the classical methanogenesis pathways, such as

211 tetrahydromethanopterin S-methyltransferase (Mtr), F420-reducing hydrogenase and Ftr, appear to have

212 been lost along the branch leading to the common ancestor of Halobacteria and

213 "Methanonatronarchaeia". After the divergence, Halobacteria continued to lose all other genes

214 involved in methanogenesis and acquire genes for aerobic and mostly heterotrophic pathways, whereas

215 "Methanonatronarchaeia" retained most pathways for anaerobic metabolism, while rewiring the

216 methanogenic pathways for the mixotrophic lifestyle (Figure 5 a). As in other cases, genome

217 reduction in "Methanonatronarchaeia" affected RNA modification, DNA repair and stress response

218 systems as well as surface protein structures 39. The subsequent gene loss occurred differentially in the

219 two groups of "Methanonatronarchaeia", suggesting adaptation to different ecological niches. The

220 HMET group lost chemotaxis and motility genes and shows signs of adaptation to heterotrophy,

221 whereas AMET retains the ability to synthesize most cellular building blocks at the expense of

222 transporter loss. The AMET strains are motile but lost attachment pili, which are present in the vast

223 majority of the species of the Halobacteria-Methanomicrobia clade 40, and many glycosyltransferases,

224 suggesting simplification of the surface protein structures. The presence of two complete CRISPR-Cas

225 systems in HMET1 compared to none in AMET1, along with the large excess of genes implicated in

226 anti-parasite defense and transposons in HMET1 (Figure 5 c and Table 1), further emphasize the

227 lifestyle differences indicating that HMET1 is subject to a much stronger pressure from mobile

228 elements than AMET1.

229

230 Central metabolism reconstruction

231 In agreement with the experimental results, genome analysis allowed us to identify the genes

232 of AMET1 and HMET1 that are implicated in energy flow and key reactions of biomass production,

233 which appear to be simple and straightforward (Figure 6). The main path starts with utilization of C1

234 methyl-containing compounds for methane production by CoM methyltransferases and methyl-CoM

235 reductase complexes, respectively. Similar to the methyl-reducing Methanomasillicoccales6, the

236 genomes of AMET1 and, especially, HMET1 contains multiple operons encoding diverse

9

237 methyltransferases (Supplementary Figure 7). Methyl-reduction is coupled with ATP generation and

238 involves five membrane-associated complexes, namely, formate dehydrogenase, membrane-bound

239 heterodisulfide reductase HdrED, Ni,Fe hydrogenase I, multisubunit Na+/H+ antiporter and H+ -

240 transporting ATP synthase. The recently characterized complete biosynthetic pathway41 for coenzyme

241 F430 is present in both genomes. In addition, membrane b-type cytochromes and methanophenazine-

242 like compounds are implicated in electron transport.

243 Pyruvate, the key entry point for biomass production, is generated through acetate

244 incorporation by acetyl-CoA synthetase (Figure 6). In a sharp contrast to most methanogens, both

245 genomes lack genes for tetrahydromethanopterin S-methyltransferase Mtr complex and

246 formylmethanofuran dehydrogenase Fwd complex, leaving all intermediate reactions, for which the

247 genes are present, unconnected to other pathways (Figure 6). All four recently reported deep lineages

248 of euryarchaeal methyl-reducing methanogens (Methanomasillilicoccales and 'Candidatus

249 Methanofastidiosa') and those from the TACK superphylum ('Candidatus Bathyarchaeota' and

250 ('Candidatus Verstraetearchaeota') 8,9 lack the Mtr and Fwd complexes as well, but they also lack all

251 the genes involved in intermediate reactions. It is extremely unlikely that genes for all Mtr and Fdw

252 complex subunits are present in both AMET1 and HMET1 but were missed by sequencing. Thus,

253 these organisms might possess still unknown pathways to connect the intermediate reactions to the rest

254 of the metabolic network.

255 In addition to the main biosynthetic pathway, AMET1 and HMET1 possess genes for three

256 key reactions of anaplerotic CO2 fixation, namely, malic enzyme, phosphoenolpyruvate carboxylase

257 and carbamoylphosphate synthase. Furthermore, complete gene sets for CO2 fixation pathway through

258 archaeal RUBISCO are present in both genomes (Figure 6) 42. The great majority of the genes

259 involved in the key biosynthetic pathways for amino acids, nucleotides, cofactors and lipids also were

260 identified in both genomes and found to be highly expressed in proteomic analysis, as revealed by

261 estimating the absolute protein amount based on the exponentially modified protein abundance index

262 (emPAI) (Supplementary Table 6 and 7). Interestingly, emPAI-based abundances follow an

263 exponential distribution in which 4 proteins involved in methanogenesis are among the 10 most highly

264 expressed proteins.

10

265 HMET1 seems to be more metabolically versatile compared with AMET1, especially with

266 respect to methanogenesis as well as amino acid and sugar metabolism (Figure 5 b and 5 c). However,

267 unlike AMET1, HMET1 lacks several genes for biosynthesis, such as quinolinate synthase

268 NadA and nicotinate-nucleotide pyrophosphorylase NadC, both involved in NAD biosynthesis;

269 uroporphyrinogen-III decarboxylase HemE and protoporphyrinogen IX oxidase HemG involved in

270 heme biosynthesis, sulfopyruvate decarboxylase involved in CoM biosynthesis and cofCED genes for

271 the coenzyme F420 biosynthesis enzyme complex. This shortage of biosynthetic enzymes is consistent

272 with experimental observations on poorer growth and survival of HMET in culture compared to

273 AMET.

274

275 Adaptation to extreme salinity

276 Given that acidification of proteins is a common feature of the "salt-in" osmotic strategy, we estimated

277 isoelectric points for the proteomes of a large representative set of halophilic and non-halophilic

278 archaea and bacteria, and compared the distributions as described under Materials and Methods

279 (Supplementary Table 5). The distributions of isoelectric points in the AMET1 and HMET1

280 proteomes are similar to those of moderately halophilic archaea and bacteria, with the notable

281 exception of their closest relatives, the extremely halophilic Halobacteria, which form a distinct cloud

282 of extremely acidic proteomes (Figure 5 d). This separation indicates that the proteome acidity of

283 Halobacteria dramatically changed after the divergence from "Methanonatronarchaeia" that appear to

284 be an evolutionary intermediate on the path from methanogens to extreme . In agreement

285 with the “salt-in” osmoprotection strategy, AMET1 and HMET1 encode a variety of K+ transporters

286 (arCOG01960) but show no enrichment of transporters for known organic osmolytes, such as glycine,

287 betaine, ectoine, or glycerol, compared with other archaea (Supplementary Table 3). On more

288 general grounds, the "salt-out" strategy appears unlikely and perhaps unfeasible for extremely

289 halophilic secondary anaerobes with relatively low energy yield. Taken together, these considerations

290 suggest that the adaptation of "Methanonatronarchaeia" to the extreme salinity relies on the "salt-in"

291 strategy. Whether these organisms possess additional mechanisms for cation-binding to compensate

11

292 for the relatively low proteome acidity, remains to be determined, but it is also possible that the main

293 counter-anion, in this case, is Cl-.

294 Analysis of the AMET1 and HMET1 protein complements revealed a major expansion of the

295 UspA family of stress response proteins with likely chaperonin function that could contribute to the

296 structural stability of intracellular proteins (Supplementary Figure 8). Finally, we identified several

297 arCOGs consisting of uncharacterized membrane proteins (eg. arCOG04755, arCOG04622,

298 arCOG04619) that are specifically shared by AMET1, HMET1 and the majority of Halobacteria

299 (Supplementary Table 3). Some of these proteins contain pleckstrin homology domains, which

300 contribute to the mechanical stability of membranes in eukaryotes 43 and might play a similar role in

301 "Methanonatronarchaeia".

302 Notably, AMET1 protein expression analysis showed that the DNA/RNA-binding protein

303 Alba, an archaeal histone and one of the UspA family proteins were among the ten most abundant

304 proteins (Supplementary Table 6). These proteins contribute to RNA, DNA and protein stability and

305 might play important roles in supporting growth under extreme salinity conditions.

306

307 Implications for the origin of methanogenesis

308 In previous phylogenetic analyses of the methyl reductase complex (McrABCD)

309 subunits, the topology of the tree for these proteins generally reproduced the ribosomal protein-based

310 phylogeny 13,22. In the present phylogenetic analysis that used different protein sets and methods,

311 AMET1/HMET1, Methanomasilliicoccales 13, ANME1 group 44 and 'Candidatus Methanofastidiosa'

312 (WSA2 group) 21 clustered together with high confidence (Supplementary Figure 9 and 10,

313 Supplementary Data 2, 3 and 4). This topology differs from the topology of the ribosomal protein

314 tree (Figure 4 a). This discrepancy could result from a combination of multiple horizontal transfers of

315 mcrABCD genes, differential gain and loss of paralogs, insufficient sampling of rare lineages, and

316 phylogenetic artefacts caused by variation of evolutionary rates. Indeed, we observed a complex

317 evolutionary history of McrA, including many lineage-specific duplications and losses

318 (Supplementary Figure 9).

12

319 Reconstruction of evolutionary events and mapping the methanogens onto the archaeal tree

320 suggests that the origin of methanogenesis dates back to the common ancestor of archaea, with

321 multiple, independent losses in various clades (Figure 5 a). The loss of the methanogenic pathways

322 often proceeds through intermediate stages as clearly observed both in "Methanonatronarchaeia" and

323 Methanomasilliicoccales (Figure 5 a). Comparison of the gene sets (arCOGs) enriched in different

324 groups of methanogens (Supplementary Table 3) using multidimensional scaling revealed distinct

325 patterns of gene loss in "Methanonatronarchaeia", Methanomasilliicoccales, ANME1 and 'Candidatus

326 Bathyarchaeota', in agreement with the independent gene loss scenario (Figure 5 e). Notwithstanding

327 these arguments, the possibility that 'Candidatus Bathyarchaeota' and ANME1 acquired

328 methanogenesis via HGT cannot be ruled out, relegating its origin to the common ancestor of

329 Euryarchaeota. Further sampling of diverse archaeal genomes should resolve this issue.

330

331 Conclusions

332 We discovered an unknown, deep euryarchaeal lineage of moderately thermophilic and extremely

333 halo(natrono)philic methanogens that thrive in hypersaline lakes. This group is not monophyletic with

334 the other methanogens but forms a separate, class-level lineage "Methanonatronarchaeia” that is most

335 closely related to Halobacteria. The "Methanonatronarchaeia” possess the methyl-reducing type of

336 methanogenesis, where C1-methylated compounds serve as acceptor and formate or H2 are external

337 electron donor, but differ from all other methanogens with this type of metabolism in the electron

338 transport mechanism. In contrast to all previously described halophilic methanogens,

339 "Methanonatronarchaeia” grow optimally in saturated salt brines and probably employ potassium-

340 based osmoprotection, similar to extremely halophilic archaea and Halanaerobiales. This discovery is

341 expected to have substantial impact on our understanding of biogeochemistry, and evolution

342 of the globally important microbial methanogenesis.

343

344 Methods 345 346 Samples 347 Anaerobic sediments (depth from 5 to 15 cm) and near bottom brines were obtained in hypersaline soda and salt 348 lakes in south-western Siberia (Altai region) and south Russia Volgograd region and Crimea) in July of 2013-

13

349 2015. The salt concentration varied from 100 to 400 g/l and the pH from 6.5-8 (salt lakes) to 9.8-10.5 (soda 350 lakes). In addition, sediments from Wadi al Natrun alkaline hypersaline lakes in Egypt (October 2000) and 351 alkaline hypersaline Searles Lake in California (April 2005) were used as inoculum in methanogenic enrichment 352 cultures. The details of the lake properties are given in Supplementary Table 1. The methanogenic potential 353 activity measurements followed by the mcrA analysis have been performed in 1:1 sediment-brine slurries as 354 described previously 28. 355 356 Enrichment and cultivation conditions 357 For soda lakes, the sodium carbonate-based mineral media containing 1-4 M total Na+ strongly buffered at pH 10 358 28,45 was used for enrichments. For salt lakes, the mineral medium containing 4 M NaCl and 0.1 M KCl buffered 359 with 50 mM K phosphates at pH 6.8 was employed. Both media after sterilization were supplied with 1 ml/l of 46 47 46 360 acidic and alkaline W/Se trace metal solutions, 1 ml/l of vitamin mix , 4 mM NH4Cl, 20 mg/l yeast extract 361 and 0.1 mM filter-sterilized CoM. The media were dispensed in serum bottles with butyl rubber stoppers of 362 various capacity at 50 (H2) - 80% (formate) filled volumes, made anoxic with 5 cycles of argon flushing- 363 evacuation and finally reduced by the addition of 1 mM Na2S and 1 drop/100 ml of 10% dithionite in 1 M 364 NaHCO3. H2 was added on the top of argon atmosphere at 0.5 bar overpressure, formate and methanol - at 50 365 mM, methylamines - at 10 mM, methyl- and dimethyl sulfides - at 5 mM. In case of methyamines, ammonium 366 was omitted from the basic medium. The incubation temperature varied from 30 to 65oC. Analyses of growth 367 parameters, pH-salt profiling of growth and activity of washed cells, optical and electron microscopy and 368 chemical analyses were performed as described previously 28,45. 369 370 Biomass composition 371 The presence of organic compatible solutes was tested by using HPLC and 1H-NMR after extraction from dry 372 cells with EtOH and the intracellular potassium concentration was quantified by ICP-MS. The presence of the 373 methanophenazine analogues was analyzed in aceton extract from lyophilized cells, followed by TLC separation, 374 reextraction with MeOH-chloroform mixture and MS-MS spectrometry. 375 376 Genome sequencing and assembly 377 The genomic DNA from pure and highly enriched cultures was obtained by using UltraClean Microbial 378 DNA Extraction Kit (MoBio Laboratories). The genome sequencing, assembly and automatic annotation of a 379 pure culture from soda lakes and of a metagenome from a highly enriched salt lake culture was performed by 380 BaseClear (Leiden, The Netherlands) using a combination of Illumina and PackBio platforms. Kmer 381 tetranucleotide frequency analysis was used to identify contigs that are likely belong to HMET1 (meta)genome. 382 Genome completeness has been estimated as described previously 48. 383 384 Genome annotation and sequence analysis 385 The final gene call has been combined from results by PROKKA 49 and GeneMarkS 50 pipelines. All 386 protein coding genes were assigned to most recent archaeal Clusters of Orthologous Groups, arCOGs as 387 described previously 34. Protein annotations were obtained by a combination of arCOGs and PROKKA 388 annotations and, in case of conflict, the respective protein have been manually reanalyzed using PSI-BLAST 51 389 and HHpred results 52 and their annotations were modified if necessary. The final assembled and annotated 390 genomic datasets were deposited in GenBank under accession numbers: MRZU00000000 and 391 MSDW00000000. 392 Other genomes related to the comparison described here were taken from GenBank (March 2016) and if 393 it was necessary ORFs have been predicted using GeneMarkS 50. 394 Protein sequences were aligned using MUSCLE 53. Alignments for the tree reconstruction have been 395 filtered to obtain informative position as described previously 34. Approximate maximum likelihood 396 phylogenetic trees were reconstructed using FastTree 54 and PHYML 55 methods. The PHYML program was 397 used for the phylogenetic tree reconstruction from an alignment of 51 concatenated ribosomal proteins (287 398 species, 8072 positions), with the following parameters: LG matrix, gamma distributed site rates, default 399 frequencies which were determined by PROTTEST program 56. Support values were estimated using an 400 approximate Bayesian method implemented in PhyML. For McrA, multiple alignment (145 sequences and 553 401 positions) was used for tree reconstruction using PhyML and PROTTEST as described above. 402 Two sets of genes reconstructed previously using the program COUNT39, which employs a Markov 403 chain gene birth and death model, for the ancestors of Halobacteriales/ 404 were used to infer gene gains and losses on the branches leading to Halobacteriales, and the discovered clade of 405 extremely halophilic methanogens. We considered an arCOG to be present in these two clades when the 406 respective COUNT probability was higher than 50%. Further reconstruction was done using a straightforward 407 parsimony approach as explained in detail in Supplementary Table 4.

14

408 Isoelectric points (pI) of individual proteins were calculated according to Bjellqvist et al. 57 using the pK 409 values from the EMBOSS suite 58. Genome-wide distributions of the protein pI were obtained as the probability 410 density estimates at 100 points in the 2.0 – 14.0 pH range using the Gaussian kernel method 59. Kullback-Leibler 411 divergence of the pI distributions for the pair of genomes A and B, DKL(A|B) was computed for all ordered pairs 60 412 of the set. The distance between the genomes was estimated as D(A,B) = D(B,A) = (DKL(A|B) + DKL(B|A))/2 . 413 The matrix of genome distances was projected into a two-dimensional space using the Classical 414 Multidimensional Scaling method 61,62 as implemented in the R package 63. 415 416 Proteomics 417 Proteomic analyses were conducted using the soda lake pure culture AMET1 (48ºC, MeOH+formate) 418 and the salt lake enrichment HMET1 (37ºC, TMA+H2) (Supplementary Table 6 and 7). Cell pellets were 419 dissolved in lysis buffer (8 M urea, 2 M thiourea, 5% CHAPS, 5 mM TCEP-HCl and a protease inhibitors 420 cocktail). Homogenization of the cells was achieved by ultra-sonication for 5 min on ultrasonic bath. After 421 homogenization, the lysed cells were centrifuged at 20,000×g for 10 min at 4 °C, and the supernatant containing 422 the solubilized proteins was used for LC-MS/MS experiment. All samples were precipitated by 423 methanol/chloroform method and re-suspended in a multi-chaotropic sample solution (7 M urea, 2 M thiourea, 424 100 mM TEAB; pH 7.5). Total protein concentration was determined using Pierce 660 nm protein assay 425 (Thermo). 40 µg of protein from each sample were reduced with 2 µL of 50mM Tris(2-carboxyethyl) phosphine 426 (TCEP, SCIEX), pH 8.0, at 37°C for 60 min and followed by 1 µL of 200mM cysteine-blocking reagent (methyl 427 methanethiosulfonate (MMTS, Pierce) for 10 min at room temperature. Samples were diluted up to 140 µL to 428 reduce urea concentration with 25mM TEAB. Digestions were initiated by adding 2 µg Pierce MS-grade trypsin 429 (Thermo Scientific) to each sample in a ratio 1:20 (w/w), which were then incubated at 37°C overnight on a 430 shaker. Sample digestions were evaporated to dryness in a vacuum concentrator and then desalted onto StageTip 431 C18 Pipette tips (Thermo Scientific) until the mass spectrometric analysis. 432 A 1 µg aliquot of each sample was subjected to 1D-nano LC ESI-MSMS analysis using a nano liquid 433 chromatography system (Eksigent Technologies nanoLC Ultra 1D plus, AB SCIEX, Foster City, CA) coupled to 434 high speed Triple TOF 5600 mass spectrometer (SCIEX, Foster City, CA) with a Nanospray III source. The 435 analytical column used was a silica-based reversed phase Acquity UPLC M-Class Peptide BEH C18 Column, 75 436 µm × 150 mm, 1.7 µm particle size and 130 Å pore size (Waters). The trap column was a C18 Acclaim 437 PepMapTM 100 (Thermo Scientific), 100 µm × 2 cm, 5 µm particle diameter, 100 Å pore size, switched on-line 438 with the analytical column. The loading pump delivered a solution of 0.1% formic acid in water at 2 µl/min. The 439 nano-pump provided a flow-rate of 250 nl/min and was operated under gradient elution conditions. Peptides 440 were separated using a 250 minutes gradient ranging from 2% to 90% mobile phase B (mobile phase A: 2% 441 acetonitrile, 0.1% formic acid; mobile phase B: 100% acetonitrile, 0.1% formic acid). Injection volume was 5 µl. 442 Data acquisition was performed with a TripleTOF 5600 System (SCIEX, Foster City, CA). Data was 443 acquired using an ionspray voltage floating (ISVF) 2300 V, curtain gas (CUR) 35, interface heater temperature 444 (IHT) 150, ion source gas 1 (GS1) 25, declustering potential (DP) 100 V. All data was acquired using 445 information-dependent acquisition (IDA) mode with Analyst TF 1.7 software (SCIEX, Foster City, CA). For 446 IDA parameters, 0.25s MS survey scan in the mass range of 350–1250 Da were followed by 35 MS/MS scans of 447 100ms in the mass range of 100–1800 (total cycle time: 4 s). Switching criteria were set to ions greater than mass 448 to charge ratio (m/z) 350 and smaller than m/z 1250 with charge state of 2–5 and an abundance threshold of 449 more than 90 counts (cps). Former target ions were excluded for 15s. IDA rolling collision energy (CE) 450 parameters script was used for automatically controlling the CE. 451 MS and MS/MS data obtained for individual samples were processed using Analyst® TF 1.7 Software 452 (SCIEX). The reconstituted AMET1 and HMET1 chromosome sequence was used to generate the database for 453 protein identification using the Mascot Server v. 2.5.1 (Matrix Science, London, UK). Search parameters were 454 set as follows: carbamidomethyl (C) as fixed modification and acetyl (Protein N-term), Gln to pyro-Glu (N-term 455 Q), Glu to pyro-Glu (N-term E) and methionine oxidation as variable modifications. Peptide mass tolerance was 456 set to 25 ppm and 0.05 Da for fragment masses, also 2 missed cleavages were allowed. The confidence interval 457 for protein identification was set to ≥ 95% (p<0.05) and only peptides with an individual ion score above the 1% 458 False Discovery Rates (FDR) at PSM level were considered correctly identified. False Discovery Rates were 459 manually calculated. The threshold of only one identified peptide per protein identification was used because

460 FDR controlled experiments counter intuitively suffer from the two‐peptide rule 64. To rank the protein

461 abundance in each sample, the Exponentially Modified Protein Abundance Index (emPAI) was used in the 462 present study as a relative quantitation score of the proteins in a complex mixture based on protein coverage by 463 the peptide matches in a database search result 62. Although the emPAI is not as accurate as quantification using 464 synthesized peptide standards, it is quite useful for obtaining a broad overview of proteome profiles. 15

465 466 Acknowledgements. DYS was supported by STW (project 12226), Gravitation-SIAM Program (grant 467 24002002, Dutch Ministry of Education and Science) and by RFBR (grant 16-04-00035). KSM, YIW and EVK 468 are supported by the intramural program of the U.S. Department of Health and Human Services (to the National 469 Library of Medicine). The proteomic analysis was performed in the Proteomics Facility of The Spanish National 470 Center for Biotechnology (CNB-CSIC) that belongs to ProteoRed, PRB2-ISCIII, supported by grant PT13/0001. 471 This project has received funding from the European Union’s Horizon 2020 research and innovation program 472 [Blue Growth: Unlocking the potential of Seas and Oceans] under grant agreement No [634486]. This work was 473 further funded by grant BIO2014-54494-R from the Spanish Ministry of Economy, Industry and 474 Competitiveness. 475 476 Author contributions. D.Y.S. performed the field work, the sediment activity incubations, enrichment and 477 isolation of pure cultures and microbiological investigation of enriched and pure cultures. B.A. and A.Y.M. 478 analyzed the mcrA and 16S rRNA gens in sediments and methanogenic cultures. M.F., P.N.G., S.C. and M.C.M. 479 were responsible for the proteomic analysis. E.G. analyzed compatible solutes. K.S.M., Y.I.W. and E.V.K. 480 performed genomic analysis and evolutionary reconstructions. D.Y.S., K.S.M. and E.V.K. wrote the paper. 481 M.C.M.L. oversaw the project and participated in the data interpretation and discussion. 482 483 Competing interests. The authors declare no competing financial interests.

16

484 References

485 1 Ferry, J. G. & Kastead, K. A. in Archaea: Molecular and Cellular Biology (ed R. 486 Cavicchioli) 288-214 (ASM Press, 2007). 487 2 Conrad, R. The global methane cycle: recent advances in understanding the microbial 488 processes involved. Environmental microbiology reports 1, 285-292, 489 doi:10.1111/j.1758-2229.2009.00038.x (2009). 490 3 Agency, U. S. E. P. (ed US-EPA) (2016). 491 4 Garrity, G. M. & Holt, J. G. in Bergey's Manual of Systematics of Archaea and 492 Bacteria Vol. 1 (John Wiley & Sons, Inc. , 2015). 493 5 Iino, T. et al. Candidatus Methanogranum caenicola: a novel methanogen from the 494 anaerobic digested sludge, and proposal of Methanomassiliicoccaceae fam. nov. and 495 Methanomassiliicoccales ord. nov., for a methanogenic lineage of the class 496 Thermoplasmata. Microbes and environments 28, 244-250 (2013). 497 6 Borrel, G. et al. Comparative genomics highlights the unique biology of 498 Methanomassiliicoccales, a -related seventh order of methanogenic 499 archaea that encodes . BMC genomics 15, 679, doi:10.1186/1471-2164-15- 500 679 (2014). 501 7 Lang, K. et al. New mode of energy metabolism in the seventh order of methanogens 502 as revealed by comparative genome analysis of "Candidatus methanoplasma 503 termitum". Applied and environmental microbiology 81, 1338-1352, 504 doi:10.1128/AEM.03389-14 (2015). 505 8 Evans, P. N. et al. Methane metabolism in the archaeal phylum Bathyarchaeota 506 revealed by genome-centric metagenomics. Science 350, 434-438, 507 doi:10.1126/science.aac7745 (2015). 508 9 Vanwonterghem, I. et al. Methylotrophic methanogenesis discovered in the archaeal 509 phylum Verstraetearchaeota. Nature microbiology 1, 16170, 510 doi:10.1038/nmicrobiol.2016.170 (2016). 511 10 Hedderich, R. & Whitman, W. B. in The – Prokaryotic Physiology and 512 Biochemistry (ed E. Rosenberg) 636-663 (Springer-Verlag, 2013). 513 11 Liu, Y. & Whitman, W. B. Metabolic, phylogenetic, and ecological diversity of the 514 methanogenic archaea. Annals of the New York Academy of Sciences 1125, 171-189, 515 doi:10.1196/annals.1419.019 (2008). 516 12 Thauer, R. K., Kaster, A. K., Seedorf, H., Buckel, W. & Hedderich, R. Methanogenic 517 archaea: ecologically relevant differences in energy conservation. Nature reviews. 518 Microbiology 6, 579-591, doi:10.1038/nrmicro1931 (2008). 519 13 Borrel, G. et al. Phylogenomic data support a seventh order of Methylotrophic 520 methanogens and provide insights into the evolution of Methanogenesis. Genome 521 biology and evolution 5, 1769-1780, doi:10.1093/gbe/evt128 (2013). 522 14 Dridi, B., Fardeau, M. L., Ollivier, B., Raoult, D. & Drancourt, M. 523 Methanomassiliicoccus luminyensis gen. nov., sp. nov., a methanogenic archaeon 524 isolated from human faeces. International journal of systematic and evolutionary 525 microbiology 62, 1902-1907, doi:10.1099/ijs.0.033712-0 (2012). 526 15 Paul, K., Nonoh, J. O., Mikulski, L. & Brune, A. "Methanoplasmatales," 527 Thermoplasmatales-related archaea in termite guts and other environments, are the 528 seventh order of methanogens. Applied and environmental microbiology 78, 8245- 529 8253, doi:10.1128/AEM.02193-12 (2012). 530 16 Fricke, W. F. et al. The genome sequence of Methanosphaera stadtmanae reveals why 531 this human intestinal archaeon is restricted to methanol and H2 for methane formation

17

532 and ATP synthesis. Journal of bacteriology 188, 642-658, doi:10.1128/JB.188.2.642- 533 658.2006 (2006). 534 17 Miller, T. L. & Wolin, M. J. Methanosphaera stadtmaniae gen. nov., sp. nov.: a 535 species that forms methane by reducing methanol with hydrogen. Archives of 536 microbiology 141, 116-122 (1985). 537 18 Sprenger, W. W., Hackstein, J. H. & Keltjens, J. T. The energy metabolism of 538 Methanomicrococcus blatticola: physiological and biochemical aspects. Antonie van 539 Leeuwenhoek 87, 289-299, doi:10.1007/s10482-004-5941-5 (2005). 540 19 Sprenger, W. W., Hackstein, J. H. & Keltjens, J. T. The competitive success of 541 Methanomicrococcus blatticola, a dominant methylotrophic methanogen in the 542 cockroach hindgut, is supported by high substrate affinities and favorable 543 thermodynamics. FEMS microbiology ecology 60, 266-275, doi:10.1111/j.1574- 544 6941.2007.00287.x (2007). 545 20 Sprenger, W. W., van Belzen, M. C., Rosenberg, J., Hackstein, J. H. & Keltjens, J. T. 546 Methanomicrococcus blatticola gen. nov., sp. nov., a methanol- and - 547 reducing methanogen from the hindgut of the cockroach Periplaneta americana. 548 International journal of systematic and evolutionary microbiology 50 Pt 6, 1989- 549 1999, doi:10.1099/00207713-50-6-1989 (2000). 550 21 Nobu, M. K., Narihiro, T., Kuroda, K., Mei, R. & Liu, W. T. Chasing the elusive 551 Euryarchaeota class WSA2: genomes reveal a uniquely fastidious methyl-reducing 552 methanogen. The ISME journal 10, 2478-2487, doi:10.1038/ismej.2016.33 (2016). 553 22 Borrel, G., Adam, P. S. & Gribaldo, S. Methanogenesis and the Wood-Ljungdahl 554 Pathway: An Ancient, Versatile, and Fragile Association. Genome biology and 555 evolution 8, 1706-1711, doi:10.1093/gbe/evw114 (2016). 556 23 McGenity, T. J. in Handbook of Hydrocarbon and Lipid Microbiology (ed K.N. 557 Timmis) 665-679 (Springer-Verlag, 2010). 558 24 Kelley, C. A., Poole, J. A., Tazaz, A. M., Chanton, J. P. & Bebout, B. M. Substrate 559 limitation for methanogenesis in hypersaline environments. Astrobiology 12, 89-97, 560 doi:10.1089/ast.2011.0703 (2012). 561 25 Oremland, R. S. & King, G. M. in Microbial mats. Physiological ecology of benthic 562 microbial communities (eds Y. Cohen & E. Rosenberg) 180-190 (American Society 563 for Microbiology, 1989). 564 26 Martin, D. D., Ciulla, R. A. & Roberts, M. F. Osmoadaptation in archaea. Applied and 565 environmental microbiology 65, 1815-1825 (1999). 566 27 Menaia, J. A. G. F. Osmotics of halophilic methanogenic archaeobacteria. Scholar 567 Archive paper (1992). 568 28 Sorokin, D. Y. et al. Methanogenesis at extremely haloalkaline conditions in the soda 569 lakes of Kulunda Steppe (Altai, Russia). FEMS microbiology ecology 91, 570 doi:10.1093/femsec/fiv016 (2015). 571 29 Ginzburg, M., Sachs, L. & Ginzburg, B. Z. Ion metabolism in a . I. 572 Influence of age of culture on intracellular concentrations. The Journal of general 573 physiology 55, 187-207 (1970). 574 30 Elevi Bardavid, R. & Oren, A. The amino acid composition of proteins from anaerobic 575 halophilic bacteria of the order Halanaerobiales. : life under extreme 576 conditions 16, 567-572, doi:10.1007/s00792-012-0455-y (2012). 577 31 Oren, A. Life at high salt concentrations, intracellular KCl concentrations, and acidic 578 proteomes. Frontiers in microbiology 4, 315, doi:10.3389/fmicb.2013.00315 (2013). 579 32 Sorokin, D. Y., Banciu, H. L. & Muyzer, G. Functional microbiology of soda lakes. 580 Current opinion in microbiology 25, 88-96, doi:10.1016/j.mib.2015.05.004 (2015).

18

581 33 Abken, H. J. et al. Isolation and characterization of methanophenazine and function of 582 phenazines in membrane-bound electron transport of Methanosarcina mazei Go1. 583 Journal of bacteriology 180, 2027-2032 (1998). 584 34 Makarova, K. S., Wolf, Y. I. & Koonin, E. V. Archaeal Clusters of Orthologous Genes 585 (arCOGs): An Update and Application for Analysis of Shared Features between 586 , , and . Life (Basel) 5, 818-840, 587 doi:life5010818 [pii] 10.3390/life5010818 (2015). 588 35 Yutin, N., Puigbo, P., Koonin, E. V. & Wolf, Y. I. Phylogenomics of prokaryotic 589 ribosomal proteins. PLoS One 7, e36972, doi:10.1371/journal.pone.0036972 PONE- 590 D-11-23203 [pii] (2012). 591 36 Eder, W., Schmidt, M., Koch, M., Garbe-Schonberg, D. & Huber, R. Prokaryotic 592 phylogenetic diversity and corresponding geochemical data of the brine-seawater 593 interface of the Shaban Deep, Red Sea. Environmental microbiology 4, 758-763 594 (2002). 595 37 Jiang, H. et al. Microbial response to salinity change in Lake Chaka, a hypersaline 596 lake on Tibetan plateau. Environmental microbiology 9, 2603-2621, 597 doi:10.1111/j.1462-2920.2007.01377.x (2007). 598 38 Yarza, P. et al. Uniting the classification of cultured and uncultured bacteria and 599 archaea using 16S rRNA gene sequences. Nature reviews. Microbiology 12, 635-645, 600 doi:10.1038/nrmicro3330 (2014). 601 39 Wolf, Y. I., Makarova, K. S., Yutin, N. & Koonin, E. V. Updated clusters of 602 orthologous genes for Archaea: a complex ancestor of the Archaea and the byways of 603 . Biol Direct 7, 46, doi:10.1186/1745-6150-7-46 1745-6150-7- 604 46 [pii] (2012). 605 40 Makarova, K. S., Koonin, E. V. & Albers, S. V. Diversity and Evolution of Type IV 606 pili Systems in Archaea. Frontiers in microbiology 7, 667, 607 doi:10.3389/fmicb.2016.00667 (2016). 608 41 Zheng, K., Ngo, P. D., Owens, V. L., Yang, X. P. & Mansoorabadi, S. O. The 609 biosynthetic pathway of coenzyme F430 in methanogenic and methanotrophic 610 archaea. Science 354, 339-342, doi:10.1126/science.aag2947 (2016). 611 42 Aono, R. et al. Enzymatic characterization of AMP phosphorylase and ribose-1,5- 612 bisphosphate isomerase functioning in an archaeal AMP metabolic pathway. Journal 613 of bacteriology 194, 6847-6855, doi:10.1128/JB.01335-12 (2012). 614 43 Baines, A. J. Evolution of spectrin function in cytoskeletal and membrane networks. 615 Biochemical Society transactions 37, 796-803, doi:10.1042/BST0370796 (2009). 616 44 Hallam, S. J., Girguis, P. R., Preston, C. M., Richardson, P. M. & DeLong, E. F. 617 Identification of methyl coenzyme M reductase A (mcrA) genes associated with 618 methane-oxidizing archaea. Applied and environmental microbiology 69, 5483-5491 619 (2003). 620 45 Sorokin, D. Y. et al. Methanosalsum natronophilum sp. nov., and 621 alkaliphilus sp. nov., haloalkaliphilic methanogens from hypersaline soda lakes. 622 International journal of systematic and evolutionary microbiology 65, 3739-3745, 623 doi:10.1099/ijsem.0.000488 (2015). 624 46 Pfennig, N. & Lippert, K. D. Über das Vitamin B12-Bedürfnis phototropher 625 Schwefelbakterien. Arch. Mikrobiol. 55, 245-256 (1966). 626 47 Plugge, C. M. Anoxic media design, preparation, and considerations. Methods in 627 enzymology 397, 3-16, doi:10.1016/S0076-6879(05)97001-8 (2005). 628 48 Podar, M. et al. Insights into archaeal evolution and symbiosis from the genomes of a 629 nanoarchaeon and its inferred crenarchaeal host from Obsidian Pool, Yellowstone 630 National Park. Biology direct 8, 9, doi:10.1186/1745-6150-8-9 (2013).

19

631 49 Seemann, T. Prokka: rapid prokaryotic genome annotation. Bioinformatics 30, 2068- 632 2069, doi:10.1093/bioinformatics/btu153 (2014). 633 50 Besemer, J., Lomsadze, A. & Borodovsky, M. GeneMarkS: a self-training method for 634 prediction of gene starts in microbial genomes. Implications for finding sequence 635 motifs in regulatory regions. Nucleic Acids Res 29, 2607-2618 (2001). 636 51 Altschul, S. F. et al. Gapped BLAST and PSI-BLAST: a new generation of protein 637 database search programs. Nucleic Acids Res 25, 3389-3402 (1997). 638 52 Soding, J., Biegert, A. & Lupas, A. N. The HHpred interactive server for protein 639 homology detection and structure prediction. Nucleic Acids Res 33, W244-248, 640 doi:10.1093/nar/gki408 (2005). 641 53 Edgar, R. C. MUSCLE: multiple sequence alignment with high accuracy and high 642 throughput. Nucleic Acids Res 32, 1792-1797 (2004). 643 54 Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2--approximately maximum- 644 likelihood trees for large alignments. PLoS One 5, e9490, 645 doi:10.1371/journal.pone.0009490 (2010). 646 55 Guindon, S. et al. New algorithms and methods to estimate maximum-likelihood 647 phylogenies: assessing the performance of PhyML 3.0. Systematic biology 59, 307- 648 321, doi:10.1093/sysbio/syq010 (2010). 649 56 Darriba, D., Taboada, G. L., Doallo, R. & Posada, D. ProtTest 3: fast selection of best- 650 fit models of protein evolution. Bioinformatics 27, 1164-1165, 651 doi:10.1093/bioinformatics/btr088 (2011). 652 57 Bjellqvist, B. et al. The focusing positions of polypeptides in immobilized pH 653 gradients can be predicted from their amino acid sequences. Electrophoresis 14, 1023- 654 1031 (1993). 655 58 Rice, P., Longden, I. & Bleasby, A. EMBOSS: the European Molecular Biology Open 656 Software Suite. Trends in genetics : TIG 16, 276-277 (2000). 657 59 Parzen, E. On Estimation of a Probability Density Function and Mode. Ann. Math. 658 Statist. 33, 1065-1076 (1962). 659 60 Kullback, S. & Leibler, R. A. On information and sufficiency. Ann. Math. Stat. 22, 79- 660 86 (1951). 661 61 Gower, J. C. Some distance properties of latent root and vector methods used in 662 multivariate analysis. Biometrika 53 (1966). 663 62 Torgeson, W. S. Theory and Methods of Scaling. (Wiley 1958). 664 63 Team, R. C. (R Foundation for Statistical Computing, Vienna, Austria, 2013). 665 64 Gupta, N. & Pevzner, P. A. False discovery rates of protein identifications: a strike 666 against the two-peptide rule. Journal of proteome research 8, 4173-4181, 667 doi:10.1021/pr9004794 (2009). 668 669

20

670 Figures legends 671 672 Fig. 1 Cell morphology of the methyl-reducing methanogens from hypersaline soda (strain 673 AMET1, a-d) and salt (strain HMET1, e-f) lakes. a - phase contrast image; b and e - total 674 electron microscopy images; c-d and f - electron microscopy images of thin sectioned cells. N 675 - nucleoide, PHA? - a possible PHA storage granule; CPM - cytoplasmic membrane; ICPM - 676 cell membrane invaginations; CW - cell wall. 677 678 Fig. 2 Growth and activity of methyl-reducing methanogens from hypersaline soda lakes. a - 679 growth dynamics of strain AMET1 with MeOH+formate at 4 M total Na+, pH 9.5 and 50oC ( -1 680 Ymax=1.5 mg protein/mM MeOH; μmax=0.012-0.015 h ). b - methanogenic activity of washed 681 cells of strain AMET1 grown with MeOH+formate (at 4 M total Na+, pH 9.5 and 48oC) with 682 various methylated e-acceptors. c - influence of temperature on growth and activity of washed 683 cells of various AMET strains at 4 M total Na+, pH 9.5 with MeOH+formate as substrate. d - 684 influence of pH at 4 M Na+ on growth and activity of washed cells of strain AMET1 at 48oC 685 with MeOH+formate as substrate. e - influence of salinity at pH 9.5 on growth and activity of 686 washed cells of strain AMET1 with MeOH+formate as substrate. In all experiments, 100 μM

687 hydrotroilite (FeS x nH2O) was added. Neither growth not activity were observed with a

688 single substrate (i.e. methylated compounds, H2 or formate alone). VCH4 is a rate of methane 689 formation, normalized either per culture volume in growth experiments or per biomass in cell 690 suspension experiments. The results represent mean values from 2 parallel experiments in 2a, 691 2b and 2d, and from a single experiment in 2c and 2e. 692

693 Fig. 3 Effect of hydrotroilite (FeS x nH2O) on growth and methanogenic activity of 694 washed and exhausted cells of the AMET1 strain 695 Growth and incubation conditions: 4 M total Na+, pH 9.8, 48oC. Substrate: 50 mM

696 CH3OH+50 mM formate. The culture was grown in the presence of sterile sand. The FeS- 697 exhausted, washed cells were obtained by prolonged incubation with substrates without 698 addition of FeS followed by washing and resuspension in a fresh buffer. The results represent 699 mean values from 3 parallel experiments. 700 701 Fig. 4 Phylogenetic analysis of “Methanonatronarchaeia" (AMET1 and HMET1) 702 The tree represents a phylogeny of archaea based on an alignment of concatenated ribosomal 703 proteins. Methanogen clades are shown in blue and Halobacteriales in orange. The inferred

21

704 methanogenic branches are highlighted in blue, the inferred loss of methanogenesis is 705 indicated by dashed red branches. The arrow indicates the likely archaeal root. All branches 706 are supported at 100% level. The original tree is available in Supplementary Data 1. 707 708 Fig. 5 Comparative genomic analysis and reconstruction of gene losses and gains 709 a. Reconstruction of gene loss and gain along in "Methanonatronarchaeia" (AMET1 and 710 HMET1) and Halobacteriales. Light blue: arCOG complement; green: gains; dark blue: 711 losses. 712 b. arCOG composition of AMET1 and HMET1. 713 c. Distribution of the differences in the arCOG composition of AMET1 and HMET1 by 714 functional categories. For each category, the number of arCOGs unique to AMET1 and 715 HMET1 is indicated. The functional classification of the COGs is described at 716 ftp://ftp.ncbi.nih.gov/pub/wolf/COGs/arCOG/funclass.tab 717 d. Multidimensional scaling analysis of isoelectric point distributions. Orange: Halobacteria; 718 green: halophilic archaea and bacteria; blue: other archaea and bacteria. 719 e. Multidimensional scaling analysis of genes enriched in methanogens. 720 721 Fig. 6 Reconstruction of the central metabolic pathways shared by 722 "Methanonatronarchaeia" 723 The main methyl-reducing pathway is shown by thick magenta arrows. Metabolically fixed

724 low molecular weight compounds are shown in green. Either gene name or respective arCOG

725 number is shown for each reaction and shown in red (details are available at Supplementary

726 Table 3). Final biosynthetic products are shown as follows: light blue for amino acids, pale

727 yellow for nucleotides, brown for lipid components, pink for cofactors. Abbreviations: MF,

728 methanofuran; H4MPT, tetrahydromethanopterin; CoM, coenzyme M, CoB – coenzyme B,

729 CoA – coenzyme A.

22

730 Table 1. Summary statistics for the AMET1 and HMET1 genomes.

731 AMET1 HMET1

Number of contigs 8 4

Total length (base pairs) 1513137 2141311

Number of coding sequences 1514 2168

GC content 38% 35%

rRNAs 5S, 16S, 23S 5S, 16S, 23S

tRNAs (for different amino acids including 31 (21) 37 (21)

pyrrolysine)

Proteins assigned to arCOGs 88% 79%

Completeness based on archaeal core arCOGs 99%# 99%#

CRISPR arrays 0 4

CRISPR-cas system subtypes - I-D, III-B

Transposon-related genes 4* 121*

Integrated elements (His2-like viruses) 3 2

732 *- Some are probably pseudogenes; # - based on 218 core arCOGs

23 ab c d

CPMCPM N CW S -layer

e f

Fig. 1 biomass CH4 MeOH formate 120 0.2 50 0.18 45

) 100 0.16 40 590 0.14 35 80 0.12 30 0.1 25 60 0.08 20 (% of maximum) (%

0.06 15 4

, formate, MeOH (mM) 40 4 0.04 10 biomass growth (OD growth biomass VCH CH 0.02 5 20 0 0 0 20 40 60 80 100 120 140 160 180 200 0 Incubation time (h) formate MeOH MeOH+f MA+f DMA+f TMA+f DMS+f a b

AMET1(growth) AMET1(activity) AMET2(growth) AMET2(activity) growth activity AMET6-2(activity) AMET8(activity) AMET-Sl(activity) AMET10(growth) 120 100

90 100 80 70 80 60 50 60 40 (% of maximum) of (% (% of maximum) of (% 4 40 4 30 VCH VCH 20 20 10 0 0 30 35 40 45 50 55 60 65 7.5 8 8.5 9 9.5 10 10.5 11

o c temperature ( C) d final pH

growth(37oC) growth(50oC) activity(37oC) activity(50oC) 100 90 80 70 60 50 40 (% of maximum) of (% 4 30

VCH 20 10 0 11.522.533.544.55

+ e Total Na (M)

Fig.2 growth activity of washed cells 1.6 800

1.4 700

1.2 600 /d) 4 1 500 /(mg protein h) 0.8 400 4

0.6 300

growth (mM CH 0.4 200

0.2 100 activity,nmol CH activity,nmol

0 0 0 100 200 300 400 500 600

FeS (mM)

Fig.3 100 x|BBBG01000001.1 408| cariaci JCM 10550 GCA 001315945.1 99 1|Mpet 2415| petrolearia DSM 11571 GCA 000147875.1 A 99 B x|JOMF01000006.1 122| mobile BP GCA 000711215.1 99 x|Metlim 2882| limicola DSM 2279 GCA 000243255.1 1|Metfor 0874| formicica SMSP GCA 000327485.1 1|MMAB1 1952| MAB1 GCA 900036045.1 100100 1|BN140 1521|Methanoculleus bourgensis MS2 GCA 000304355.2 AMET1 55 99 2|MMAB1 2235|Methanoculleus MAB1 GCA 900036045.1 Methanonatronarchaeia 2|BN140 1738|Methanoculleus bourgensis MS2 GCA 000304355.2 100 100100 1|Memar 0613|Methanoculleus marisnigri JR1 GCA 000015825.1 HMET1 2|Memar 0378|Methanoculleus marisnigri JR1 GCA 000015825.1

89 1|Metli 1710| liminatans DSM 4140 GCA 000275865.1 88 1|BCNW01000001.1 1265|Methanofollis ethanolicus HASU GCA 001571385.1 99 2|Mpet 1991|Methanolacinia petrolearia DSM 11571 GCA 000147875.1 Methanomicrobiales 2|Metli 0498|Methanofollis liminatans DSM 4140 GCA 000275865.1 100 100 2|BCNW01000001.1 695|Methanofollis ethanolicus HASU GCA 001571385.1

100 1|Mlab 1068| labreanum Z GCA 000015765.1 100 100 2|Mlab 1561|Methanocorpusculum labreanum Z GCA 000015765.1 x|XE11 1766|Methanomicrobiales archaeon 53 19 GCA 001508695.1 Halobacteriales 99 100 0|Mhun 2148| hungatei JF 1 GCA 000013445.1 0|Mpal 2309|Methanosphaerula palustris E1 9c GCA 000021965.1 100 x|AGIY02000001.1 1682|Methanolinea tarda NOBI 1 GCA 000235685.3 2|Metfor 2609|Methanoregula formicica SMSP GCA 000327485.1 99 79 0|Mboo 0582|Methanoregula boonei 6A8 GCA 000017625.1

99 0|Mtc 0908| conradii HZ254 GCA 000251105.1 0|MCP 0516|Methanocella paludicola SANAE GCA 000011005.1 100 0|RCIX2063|Methanocella arvoryzae MRE50 GCA 000063445.1 Methanocellales

99 0|MCMEM 0559| methylutens MM1 GCA 000970325.1 68 0|Mbur 2417|Methanococcoides burtonii DSM 6242 GCA 000013725.1 Methanosarcinales 99 0|Mmah 0612|Methanohalophilus mahii DSM 5219 GCA 000025865.1 45 0|Metho 1670| hollandica DSM 15978 GCA 000328665.1 100 0|Mpsy 0828| psychrophilus R15 GCA 000306725.1 100 0|Metev 0902|Methanohalobium evestigatum Z 7303 GCA 000196655.1 0|Mzhil 0857|Methanosalsum zhilinae DSM 4017 GCA 000217995.1 0|MSTHT 1644|Methanosarcina thermophila TM 1 GCA 000969885.1 Methanocellales 100 99 0|MSTHC 1645|Methanosarcina thermophila CHTI 55 GCA 000969925.1

87 0|MSBRW 0105| Wiesmoor GCA 000969985.1 Methanomicrobiales 0|Mbar A0893|Methanosarcina barkeri Fusaro GCA 000195895.1 100 0|MSBR2 2279|Methanosarcina barkeri 227 GCA 000970065.1 100 0|MCM1 2594|Methanosarcina barkeri CM1 GCA 001027005.1 89 0|MSBRM 0101|Methanosarcina barkeri MS GCA 000970025.1 Archaeoglobales 100 0|MSBR3 2340|Methanosarcina barkeri 3 GCA 000970305.1 100 99 0|MSKOL 3374|Methanosarcina Kolksee Kolksee GCA 000969945.1 Thermoplasmatales 0|MSVAZ 3413|Methanosarcina vacuolata Z 761 GCA 000969905.1 100 0|MSMAS 0073|Methanosarcina mazei S 6 GCA 000970205.1 0|MM 1240|Methanosarcina mazei Go1 GCA 000007065.1 95 0|MSSIT 3838|Methanosarcina siciliae T4 M GCA 000970085.1 98 Methanomassiliicoccales 100 0|MSSIH 3759|Methanosarcina siciliae HI350 GCA 000970125.1 100 99 0|MSSAC 4255|Methanosarcina siciliae C2J GCA 000970145.1 Methanosarcinales 0|MA 4546|Methanosarcina acetivorans C2A GCA 000007345.1

61 99 0|MSWHS 0092|Methanosarcina WWM596 WWM596 GCA 000969965.1 98 Thermococcales 100 0|MSWH1 0089|Methanosarcina WH1 WH1 GCA 000970005.1 99 0|MSMTP 0100|Methanosarcina MTP4 MTP4 GCA 000970045.1 100 0|MSLAZ 0156|Methanosarcina lacustris Z 7289 GCA 000970265.1 0|MSMAL 0075|Methanosarcina mazei LYC GCA 000970225.1 Methanococcales 99 99 0|MSMAP 0074|Methanosarcina mazei SarPi GCA 000970185.1 0|MSMAC 0072|Methanosarcina mazei C16 GCA 000970245.1 0|MSMAW 0076|Methanosarcina mazei WWM610 GCA 000970165.1 100 Hadesarchaea archaeon YNP N21 0|MmTuc01 1281|Methanosarcina mazei Tuc01 GCA 000341715.1 0|MSHOH 0084|Methanosarcina horonobensis HB 1 JCM 15518 GCA 000970285.1 x|MPEBLZ 01201|Candidatus Methanoperedens BLZ1 GCA 001317315.1 kandleri AV19 () 100 x|ANME2D 01104|Candidatus Methanoperedens nitroreducens ANME 2d GCA 000685155.1 100 x|JONQ01000011.1 86|Methermicoccus shengliensis DSM 18856 GCA 000711905.1 x|XD62 0880|Methanosarcinales archeaon 56 1174 GCA 001508185.1 100 Methanothermobacteriales x|XD46 0340|Euryarchaeota archaeon 55 53 GCA 001507975.1 93 0|Mthe 0569| thermophila PT GCA 000014945.1 100 0|MCON 0759| GP6 GCA 000204415.1 0|Mhar 0498|Methanosaeta harundinacea 6Ac GCA 000235565.1 DPANN superphylum 100 x|AAQ63481.1|ANME2

43 0|Metin 0281| infernus ME GCA 000092305.1 100 2|JH146 1056|Methanocaldococcus bathoardescens JH146 GCA 000739065.1 99 100 2|Metvu 0149|Methanocaldococcus vulcanius M7 GCA 000024625.1 2|MJ 0846|Methanocaldococcus jannaschii DSM 2661 GCA 000091665.1 99 99 0|Mefer 0938|Methanocaldococcus fervens AG86 GCA 000023985.1 99 2|MFS40622 1315|Methanocaldococcus FS406 22 FS406 22 GCA 000025525.1 C. Caldiarchaeum subterraneum 2|Metig 1237| igneus Kol 5 GCA 000214415.1 0|Mvol 1128| voltae A3 GCA 000006175.2 100 100 0|Mevan 0872|Methanococcus vannielii SB GCA 000017165.1 Methanococcales 1 0|MmarC6 1111|Methanococcus maripaludis C6 GCA 000018485.1 100 Candidatus Bathyarchaeota 0|MmarC7 0806|Methanococcus maripaludis C7 GCA 000017225.1 93 0|MmarC5 0017|Methanococcus maripaludis C5 GCA 000016125.1 96 100 C. Korarchaeum cryptofilum OPF8 99 0|MMP1559|Methanococcus maripaludis S2 GCA 000011585.1 100 0|GYY 08645|Methanococcus maripaludis X1 GCA 000220645.1 100 0|Metok 0956| okinawensis IH1 GCA 000179575.2 99 0|Maeo 1268|Methanococcus aeolicus Nankai 3 GCA 000017185.1 1|Metig 0445|Methanotorris igneus Kol 5 GCA 000214415.1 99 1|Metvu 1194|Methanocaldococcus vulcanius M7 GCA 000024625.1 1|JH146 0503|Methanocaldococcus bathoardescens JH146 GCA 000739065.1 100 Methanococcales 2 98 1|MJ 0083|Methanocaldococcus jannaschii DSM 2661 GCA 000091665.1 99 1|MFS40622 0424|Methanocaldococcus FS406 22 FS406 22 GCA 000025525.1 1|Mfer 0734|Methanothermus fervidus DSM 2088 GCA 000166095.1 100 99 1|MTBMA c15120|Methanothermobacter marburgensis Marburg GCA 000145295.1

100 1|MTH 1129|Methanothermobacter thermautotrophicus Delta H GCA 000008645.1 100 99 1|MTCT 1022|Methanothermobacter CaT2 CaT2 GCA 000828575.1 1|Metbo 2281|Methanobacterium lacus AL 21 GCA 000191585.1 archaeon SMTZ1 83 100 100 1|MBMB1 1825|Methanobacterium MB1 GCA 000499765.1 1|MB9 2187|Methanobacterium formicicum GCA 001458655.1 100 1|DSM1535 1095|Methanobacterium formicicum GCA 000953115.1 Methanothermobacteriales 2 99 Lokiarchaeum GC14 75 99 1|BRM9 2156|Methanobacterium formicicum BRM9 GCA 000762265.1 97 0|Msp 0321|Methanosphaera stadtmanae DSM 3091 GCA 000012545.1 1|YLM1 1111|Methanobrevibacter olleyae YLM1 GCA 001563245.1 55 97 100 1|sm9 1353|Methanobrevibacter millerae SM9 GCA 001477655.1 99 1|Msm 0902|Methanobrevibacter smithii ATCC 35061 GCA 000016525.1 0.2 0|MK0655|Methanopyrus kandleri AV19 GCA 000007185.1 2|Mfer 0784|Methanothermus fervidus DSM 2088 GCA 000166095.1 Methanopyrales x|XD44 1486x|Methanobacteriaceae archaeon 41 258 GCA 001507955.1 2|Metbo 0353|Methanobacterium lacus AL 21 GCA 000191585.1 100 100 2|MBMB1 1682|Methanobacterium MB1 GCA 000499765.1 2|DSM1535 0905|Methanobacterium formicicum GCA 000953115.1 100 99 2|MB9 1989|Methanobacterium formicicum GCA 001458655.1 99 100 2|BRM9 0935|Methanobacterium formicicum BRM9 GCA 000762265.1 0|MSWAN 2056|Methanobacterium paludis SWAN1 GCA 000214725.1 0|Abm4 1526|Methanobrevibacter AbM4 AbM4 GCA 000404165.1 58 Fig.4. Phylogenetic analysis of “Methanonatronarchaeia" (AMET1 and HMET1) 100 2|YLM1 1379|Methanobrevibacter olleyae YLM1 GCA 001563245.1 100 0|mru 1924|Methanobrevibacter ruminantium M1 GCA 000024185.1 Methanothermobacteriales 1 100 100 2|Msm 1015|Methanobrevibacter smithii ATCC 35061 GCA 000016525.1

100 2|sm9 2027|Methanobrevibacter millerae SM9 GCA 001477655.1 a. Phylogeny of archaea based on alignment of concatenated ribosomal proteins. 99 0|TL18 09320|Methanobrevibacter YE315 YE315 GCA 001548675.1 2|MTH 1164|Methanothermobacter thermautotrophicus Delta H GCA 000008645.1 methanogen clades are shown in blue and Halobacteriales in orange 100 2|MTBMA c15480|Methanothermobacter marburgensis Marburg GCA 000145295.1 85 2|MTCT 1058|Methanothermobacter CaT2 CaT2 GCA 000828575.1 x|ALG76131.1|uncultured archaeon The inferred methanogenic branches are highlighted in blue, the inferred loss of 100 x|ANME1 1|ANME1 ANME-1 group methanogenesis is indicated by dashed red branches. The arrow indicates the likely 100 100 x|GZfos19C7 1|uncultured archaeon GZfos19C7 x|KYC53403.1|Arc I group archaeon U1lsi0528 Bin055 x|ADurb1213 1|Arc I group archaeon ADurb1213 Bin02801 archaeal root. All branches are supported at 100% level. The original tree is available 100 x|B15fssc0709 1|Arc I group archaeon B15fssc0709 Meth Bin003 Candidatus Methanofastidiosa 63 in Supplementary fig. S6. 96 x|KYC51830.1|Arc I group archaeon U1lsi0528 Bin089 2|AMET1 1458|AMET1 99 100 b. Phylogeny of Methyl coenzyme M reductase alpha subunit (McrA). Branch support 100 1|AMET1 1459|AMET1 Methanonatronarchaeia 0|HMET1 2442|HMET1 values calculated by PhyML are shown and for the clade containing AMET1 and HMET1 x|AOA81 01015|Methanomassiliicoccales archaeon RumEn M2 GCA 001421175.1 100 x|AUP07 0616|methanogenic archaeon mixed culture ISO4 G1 GCA 001563305.1 99 0|Mpt1 c02010|Candidatus Methanoplasma termitum MpT1 GCA 000800805.1 are colored red. The sequences are denoted by locus tag ID or genome partition ID, 100 0|TALC 00472|Thermoplasmatales archaeon BRNA1 BRNA1 GCA 000350305.1 98 0|AR505 1396|methanogenic archaeon ISO4 H5 ISO4 H5 GCA 001560915.1 Methanomassiliicoccales genome names and its code on the genbank FTP site. First part of the sequence 100 100 97 0|MMALV 04000|Candidatus Methanomethylophilus alvus Mx1201 GCA 000300255.2 name indicates the following: x- incomplete genome (colored gray), 0|MMINT 15640|Candidatus Methanomassiliicoccus intestinalis Issoire Mx1 GCA 000404225.1 91 x|LFRW01000043.1 38|Methanomicrobia archaeon DTU008 GCA 001512965.1 x|V4 1|Archaeon V4 100 0 – one McrA gene in complete genomes (colored black), 1 – first copy of McrA and x|V3 1|Archaeon V3

100 x|V2 1|Archaeon V2 Candidatus Verstraetearchaeota 2 – second copy of McrA (both colored blue). 100 x|V1 1|Archaeon V1 x|AOA66 1761|Bathyarchaeota archaeon BA2 GCA 001399795.1 The original tree is available in Supplementary fig. 9. 0.10 100 x|AOA65 0408|Bathyarchaeota archaeon BA1 GCA 001399805.1 Candidatus Bathyarchaeota Heme biosynthesis genes HemE, HemN A Carbonic anhydrase Methyl-CoB:CoM M methyltransferase Mtb, Mtt +51 Cytochrome c554 AMET11 Cytochrome c-type biogenesis genes -305 NADH dehydrogenase Nuo ABC-type oligopeptide transporter Dpp 1119 Glycosyltransferase Type IV pili +48 CRISPR-Cas system type III-B ancestral genes Methionine synthase II -459 retained in all 3 Several ferredoxins Multiheme cytochrome lineages cytochrome b556 FdnI Glycogen synthase Preprotein translocase SecD,SecF UspA family genes +76 2-phosphoglycerate kinase Excinuclease ABC Uvr HMET1 Mismatch repair MutL, MutS -139 NAD biosynthesis NadA,NadC Last ancestor of Acyl-CoA synthetase Chemotaxis 1301 (695) Succinyl-CoA synthetase Archaellum Halobacteria + 145 Nucleotide sugars metabolism and 2132 Glycosyl transferases Agl - 511 Cytochrome oxidase genes Methanomicrobia Pyruvate/2-oxoglutarate dehydrogenase complex Aerobic carbon monoxide dehydrogenase CO dehydrogenase Cdh Halocyanin Coenzyme F420-reducing hydrogenase Frh Rhodopsin Formylmethanofuran dehydrogenase Fwd Succinate dehydrogenase Heterodisulfide reductase HdrA, HdrC Urease Tetrahydromethanopterin S-methyltransferase Mtr Phosphotransferase system +1612 Gas vesicle system Halobacteriales -224 cytochrome b, FdhI methanogenesis genes Mtr, Mtd, Ftr 3154 Pyruvate:ferredoxin oxidoreductase POR Fumarate hydratase class I Glutamate synthase CRISPR-Cas system type III-B B C HMET1 80 AMET1 Ni,Fe-hydrogenase III 70 370 Chemotaxis 60 204 Archaellum 914 50

40

30 NADH dehydrogenase Methanogenesis genes ABC peptide transporters Pyrrolysine related genes 20 Many glycosyl transferases Anaerobic formate dehydrogenase 10 Type IV pili Cytochromes c and b 0 3 CRISPR-Cas systems Ni,Fe-hydrogenase I CEFGHMPIQJLODKTNVRS His2-like virus insertions AMET1 HMET

0.4 E D 0.2 Thermoplasmata methanogens 0.3 0.15 Bathyarchaeon 0.1 0.2 ANME1 0.05 0.1

0 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0 -2 -1 0 1 2 3 4 5 -0.05

-0.1 other methanogens -0.1 HMET1 -0.2 AMET1 -0.15

-0.3 -0.2 AMET1 -0.25 -0.4

-0.3 HMET1 Fig.5 Comparative genomic analysis and reconstruction of losses and gains a. Reconstruction of gene loss and gain along in "Methanonatronarchaeia" (AMET1 and HMET1) and Halobacteriales. Light blue: arCOG complement; green: gains; dark blue: losses. b. arCOG composition of AMET1 and HMET1. c. Distribution of arCOGs in AMET1 and HMET1 by COG functional categories. COG functional groups are described at ftp://ftp.ncbi.nih.gov/pub/wolf/COGs/arCOG/funclass.tab d. Multidimensional scaling analysis of isoelectric point distributions. Orange: Halobacteria; green: halophilic archaea and bacteria; blue: other archaea and bacteria e. Multidimensional scaling analysis of genes enriched in methanogens Sul1/Mot1* Wtp** Tup Pst MMT1 Mgt Mgt molibdate molibdate tungstate phosphate Co2+/Zn2+/Cd2+ Ca2+/Na+ Mg2+ cobalt or sulfate CbiM glucose-6-P D-arabino- K+ hexulose-6-P arCOG00103 Kef-type pgi HxlB tryptophane fructose-6-P dihydro-acetone-P D-ribulose-5-P D-ribose-5-P tyrosine RpiA GldA phenylalanine arCOG04180 SuhB PrsA fructose-1,6-PP tpi sn-Glycerol-1P arsenite arCOG04180 D-ribulose-1,5-P 5-Phosphoribosyl ASR3 glyceraldehyde-3-P GapA diphosphate E2B2 amino acid 1,3-PP-glycerate gor D-ribose-1,5-P PotE CO2 pgk RUBISCO serine DeoA 3-P-glycerate histidine glycine AMP purines cysteine ApgM pyrimidines alanine Adk riboflavin Cytochrome C 2-P-glycerate FAD Na+/proline leucine ADP Ccm valine folate PutP eno molybdopterin pantothenate Ndk CoA acetate Na+ or H+/ asparagine Phosphoenolpyruvate ATP threonine ActP methionine PpsA 2- NO3 leucine PpcA LldD Na+ /H+ Tau pyruvate L-lactate lysine CO2 pyrolysine CoA ATP Mnh Por formate isoleucine CO2 Acs Formate Fe2+/ Fdred glutamate arCOG04435 dehydrogenase Fdh Mn2+ Fdox AMP CO CCC1 HS-CoB 2 aceto- + cytochrome b CytB acetyl-CoA Isoprenoids Fe PaaJ acetyl-CoA HS-CoM 2+ SfcA Heterodisulfide FeoA/FeoB-like reductase + HdrED arCOG01130 CH4 H L-aspartate oxaloacetate citrate CoM-S-S-CoB red Fd Fdox Fdx, Fer arCOG04278 Fep-like mdh Mcr Ni,Fe H2 Fe cys-aconitate Hydrogenase 2+/ Nicotinamid- malate Methyl-S-CoM Mn HyaABD 2+ mononucleotide Zhu 2-oxoglutarate ATP Methyl-H4MPT FumA Lpd ADP + Mer H Ntp fumarate succinyl-CoA NAD Cobalamine Methylene-H4MPT Proline Mtd MtaA ArgG MttCB cysteine ProC GdhA Methenyl-H4MPT Arginino TMA FeS Pyrroline-5- Mch HS-CoM succinate ArgH Arginine carboxylate alanineSufCB Ni-sirohydrochlorin MtbCB L-glutamate Formyl-H4MPT putA CfbB HS-CoM 2- 2- Glutamate MFR DMA SO or PO NH HS-CoM 4 4 Citrulline 5-semialdehyde GlnA 3 Ni-sirohydro Ftr H4MPT MtmCB PitA ArgF HS-CoM ArgD chlorin Tricarboxylate CO L-Glutamine Formyl-MFR Methyl-S-CoM Ornithine 2 CfbA a,c -diamide HS-CoM MMA CfbCD Carbamoyl-P CarAB MtaCB sirohydrochlorin 15,173-seco- CfbE Methanol CysG F430 arCOG04354 siroheme F430-173-acid MtsCB Methanetiol or DMS

The main mixotrophic pathways are shown by thick magenta arrows. Metabolically fixed low molecular weight compounds are shown in green. Either gene name or respective arCOG number is shown for each reaction and shown in red (details are available at Supplementary Table S3). Final biosynthetic products are shown as follows: light blue for amino acids, pale yellow for nucleotides, brown for lipid components, pink for cofactors. Abbreviations: MF, methanofuran; H4MPT, tetrahydromethanopterin; CoM, coenzyme M, CoB – coenzyme B, CoA – coenzyme A.