Skip to main content
The ISME Journal logoLink to The ISME Journal
. 2024 Jul 29;18(1):wrae147. doi: 10.1093/ismejo/wrae147

Reversed oxidative TCA (roTCA) for carbon fixation by an Acidimicrobiia strain from a saline lake

Lei Gao 1,2,#, Lan Liu 3,#,, Ai-Ping Lv 4,#, Lin Fu 5, Zheng-Han Lian 6, Takuro Nunoura 7, Brian P Hedlund 8,9, Qing-Yu Xu 10,11, Dildar Wu 12, Jian Yang 13, Mukhtiar Ali 14, Meng-Meng Li 15, Yong-Hong Liu 16, André Antunes 17, Hong-Chen Jiang 18,19, Lei Cheng 20, Jian-Yu Jiao 21,, Wen-Jun Li 22,23,, Bao-Zhu Fang 24,
PMCID: PMC11697166  PMID: 39073917

Abstract

Acidimicrobiia are widely distributed in nature and suggested to be autotrophic via the Calvin–Benson–Bassham (CBB) cycle. However, direct evidence of chemolithoautotrophy in Acidimicrobiia is lacking. Here, we report a chemolithoautotrophic enrichment from a saline lake, and the subsequent isolation and characterization of a chemolithoautotroph, Salinilacustristhrix flava EGI L10123T, which belongs to a new Acidimicrobiia family. Although strain EGI L10123T is autotrophic, neither its genome nor Acidimicrobiia metagenome-assembled genomes from the enrichment culture encode genes necessary for the CBB cycle. Instead, genomic, transcriptomic, enzymatic, and stable-isotope probing data hinted at the activity of the reversed oxidative TCA (roTCA) coupled with the oxidation of sulfide as the electron donor. Phylogenetic analysis and ancestral character reconstructions of Acidimicrobiia suggested that the essential CBB gene rbcL was acquired through multiple horizontal gene transfer events from diverse microbial taxa. In contrast, genes responsible for sulfide- or hydrogen-dependent roTCA carbon fixation were already present in the last common ancestor of extant Acidimicrobiia. These findings imply the possibility of roTCA carbon fixation in Acidimicrobiia and the ecological importance of Acidimicrobiia. Further research in the future is necessary to confirm whether these characteristics are truly widespread across the clade.

Keywords: Acidimicrobiia, carbon fixation, roTCA, CBB cycle, chemolithoautotrophic enrichment, metagenome, transcriptome, stable-isotope probing, ancestral character reconstruction

Introduction

The class Acidimicrobiia is a deep-rooted lineage within the phylum Actinomycetota [1–6] that includes three validly published families under the International Code of Nomenclature of Prokaryotes (ICNP) (https://lpsn.dsmz.de/order/acidimicrobiales): Acidimicrobiaceae, Iamiaceae, and Ilumatobacteraceae [7, 8]. Acidimicrobiaceae are thermoacidophilic microbes capable of oxidizing/ reducing iron and sulfur compounds [3, 9, 10]. In contrast, Iamiaceae and Ilumatobacteraceae have mainly been found in marine environments and can grow under more moderate temperatures (15–55°C) as well as pH values (pH 6–11) [11, 12]. Despite the widespread occurrence of Acidimicrobiia in various habitats and some existing knowledge about them, the broad ecological characteristics of this taxon remain elusive, primarily due to the limited number and diversity of pure cultures as well as limited research focusing on their ecological functions [6, 13]. It has been reported that Actinomycetota contribute to the synthesis of organic matter through chemolithoautotrophy, with Acidimicrobiia playing a key role in this process [14–16]. Carbon fixation by chemolithoautotrophic microbes had a profound influence on the transition from the inorganic to the organic world via biological CO2 reduction, an indispensable process for sustaining life on Earth as it is accountable for the synthesis of the majority of organic carbon molecules [14, 17]. Probing the carbon fixation abilities of Acidimicrobiia is important for broadening our understanding of microbial diversity, carbon cycling, and biogeochemical processes. Bay et al. and Norris et al. reported several Acidimicrobiia metagenome-assembled genomes (MAGs) that encode genes for uptake hydrogenases and RuBisCO (including oxygen-producing photoautotrophic RuBisCO type IA and chemoautotrophic RuBisCO type IE), which suggested the potential for hydrogenotrophic carbon fixation by some Acidimicrobiia via the Calvin–Benson–Bassham (CBB) cycle [15, 16, 18, 19]. Although these studies provided valuable insights into their potential ability to fix carbon, no experimental evidence for carbon fixation in this class has been reported.

Here, we report a new isolate belonging to the class Acidimicrobiia, herein named Salinilacustristhrix flava EGI L10123T, that belongs to a new family Salinilacustritrichaceae fam. nov. We provide evidence of sulfide-dependent chemolithoautotrophy via the roTCA in this taxon. Moreover, ancestral character state reconstructions of key enzymes in Acidimicrobiia hint at the possibility of CO2 fixation via the roTCA.

Materials and methods

Sample collection and chemolithoautotrophic enrichment

Surface sediment samples from a depth of ~20 cm were collected from a hypersaline lake, Barkol Lake (92°48′11″E, 43°40′05″N), in Xinjiang Province, PR China (Fig. S1). The sediment at the sampling site exhibits a total salt content of 207 g/kg and a pH of 8.9, as analyzed under the People’s Republic of China National Standard for Environmental Quality Standard for Soils (GB 15618-1995) at the Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences. Similarly, the water at the sampling site shows a salinity of 66 g/L and a pH of 7.9, as determined under the People’s Republic of China National Standard for Environmental Quality Standard for Surface Water (GB 3838-2002) at the same institute. Two samples of Barkol Lake were collected (one in 2020 and the other in 2021). Samples for DNA extraction were stored at −80°C in sterile sampling tubes and samples for laboratory enrichment were kept at 4°C in sterile sampling bags. To ensure the sediments were homogenized as much as possible, we continuously stirred the samples overnight before inoculation. Three replicates of chemolithoautotrophic enrichment cultures were constructed (Fig. 1A) by using a chemolithoautotrophic enrichment M3 medium (+V) containing a vitamin solution, trace elements, and Na2S as an electron donor (Table S1), with the Barkol Lake sample collected in 2020. The medium was dispensed in 100 mL volumes in 250 mL serum bottles covered in a vented sealing film, inoculated with 5 g of sediment, and incubated at 37°C under oxic conditions (atmospheric air). All glassware was rinsed twice with 6 M HCl, followed by three rinses with Milli-Q water, autoclaved, and dried at 60°C before use to remove any contaminants that could interfere with the experiments as thoroughly as possible [22].

Figure 1.

Figure 1

Acidimicrobiia enrichments and diversity in diverse environments. (A) Schematic representation of the enrichment strategy for chemolithoautotrophic microorganisms in the saline lake. (B) Analysis of the community compositions of three original sediment samples from the saline lake and samples after chemolithoautotrophic enrichment. (C) Comparative assessment of the relative abundance of Acidimicrobiia pre- and post-chemolithoautotrophic enrichment (error bar represents standard deviation). (D) Phylogenetic analysis of Acidimicrobiia. The maximum-likelihood phylogenetic tree was constructed using multiple sequence alignments of 120 bacterial marker genes generated by GTDB-Tk [20] according to IQ-Tree software [21]. The best-fit model LG + F + I + I + R10 was well-supported by Bayesian Information Criterion. Bootstrap values were based on 1000 replicates and nodes with percentages >90% indicated as solid circles. Circles outside the tree labels illustrate genome sources. The genomes with star was obtained from this study.

DNA extraction, sequencing, and analysis of 16S rRNA gene amplicons

DNA for high-throughput sequencing from the original sample and enrichment cultures was extracted using the FastDNA Spin Kit (MP Biomedicals). The integrity and concentration of the extracted DNA were checked by agarose gel electrophoresis and Invitrogen Qubit Fluorometer. Samples for amplicon sequencing were sent to Beijing Biomarker Technologies Co, Ltd for PCR amplification, library construction, and sequencing. The V4 region of the 16S rRNA gene was amplified using the primers 515F (5’-GTGCCAGCMGCCGCGGTAA-3′) and 806R (5’-GGACTACHVGGGTWTCTAAT-3′) [23]. Amplicons were then sequenced using the NovaSeq platform (Illumina, USA). The raw amplicon reads were processed using the cutadapt plugin with parameters —p-front-f GTGCCAGCMGCCGCGGTAA —p-front-r GGACTACHVGGGTWTCTAAT in the QIIME2 (ver. 2023.2) platform to remove the primers [24]. Then, the DADA2 plugin in the QIIME2 platform was used to merge, denoise, filter chimeras, dereplicate, and obtain feature tables as well as representative amplicon sequence variant (ASV) sequences [25]. Each ASV was classified using the qiime feature-classifier plugin based on the SILVA database (Release 138.1) [26–28].

Metagenomic sequencing and analysis

For metagenomic sequencing, DNA samples were prepared using a VAHTS Universal Plus DNA Library Prep Kit for Illumina, and the insert size was assessed using Agilent 2100. Metagenomic sequencing was performed using the NovaSeq platform with PE150 (Illumina, USA) at Beijing Biomarker Technologies Co, Ltd. Approximately 30 Gbp of metagenomic data for each sample was generated (Table S1). The raw metagenomic reads were quality filtered to eliminate adapters, duplicated reads, and low-quality reads by using fastp (ver. 0.23.4) with parameters -q 20 -u 20 -e 20 -l 50 -5 -3 --cut_front_window_size 50 --cut_front_mean_quality 20 --cut_tail_window_size 50 --cut_tail_mean_quality 20 -w 20 [29, 30]. High-quality reads were de novo assembled into scaffolds using SPAdes (ver. 3.15.5) with parameters --meta -k 21,33,55,77,99 [31]. Genome binning was performed by using MetaBAT2 with the scaffolds >2500 bp [32]. CheckM (ver. 1.2.2) was used to analyze the completeness and contamination of the MAGs [33]. MAGs with estimated completeness <80% or contamination >5% were discarded. MAGs were taxonomically classified using GTDB-Tk (ver. 2.3.2) [20]. Protein-coding regions were predicted using Prodigal (ver. 2.6.3) [34]. MAGs were further annotated by querying the predicted coding sequence (CDS) against the Clusters of Orthologous Groups (COG) [35] as well as Kyoto Encyclopedia of Genes and Genomes (KEGG) databases (Release 108.1) using DIAMOND (ver. 0.7.9; E-values <1e-5) [36, 37].

Isolation and identification of a new chemolithoautotrophic Acidimicrobiia family

We employed a culture-dependent method [38, 39] to selectively isolate Acidimicrobiia from both the enrichment culture and the original sediment samples using nutrient medium (0.5 × Marine Agar 2216E, pH = 7.6, BD Difco™) as well as the chemolithoautotrophic M3 medium (+V, pH = 7.8) supplemented with 20% agar under oxic conditions (atmospheric air). For sediments, 1 g was added to 9 mL of sterile water, diluted to 10−3 in sterile water, and then spread onto plates for isolation. For the enrichment sample, 1 mL enrichment sample was added to 9 mL of sterile water, diluted to 10−3 in sterile water, and then spread onto plates for isolation. Pure cultures were obtained by repeating streaking of isolated colonies. Isolates were stored in a 20% glycerol solution at −80°C [39] and identified by amplifying near-complete 16S rRNA genes using primers (27F: 5’-AGAGTTTGATCCTGGCTCAG-3′ and 1492R: 5’-TACGACTTAACCCCAATCGC-3′) [40] followed by sequencing at Sangon Biotech (Shanghai) Co, Ltd. 16S rRNA gene sequences were analyzed using EzBioCloud (Release Note 20230823, https://www.ezbiocloud.net/). From these efforts, a single new isolate belonging to Acidimicrobiia was successfully obtained. This isolate has been officially designated as strain EGI L10123T. A polyphasic taxonomy approach [41] was conducted as described in the Supplementary Information. To assess chemolithoautotrophic growth of strain EGI L10123T under oxic conditions, a 1 mL cell suspension was inoculated into 250 mL sterile conical flasks under atmospheric air, each equipped with a vented sealing film, and filled with 100 mL of M3 medium (+V) or 100 mL of M3 medium deficient in vitamins (-V) using bicarbonate as the carbon source (Table S1), sulfide as the electron donor, and oxygen (from air) as an electron acceptor. To assess the chemolithoautotrophic growth potential of strain EGI L10123T under anoxic conditions with N2 headspace gas, a 1 mL cell suspension was inoculated into 250 mL sterile serum bottles and filled with 100 mL of M3 medium (+V) or 100 mL of M3 medium deficient in vitamins (-V) using bicarbonate as the carbon source (Table S1). After consecutive transfers in M3 medium (-V), the number of cells was measured by flow cytometry to assess chemolithoautotrophic growth.

We also quantitatively assessed the enzymatic activity of citrate synthase (CS) in strain EGI L10123T under chemolithoautotrophic conditions in M3 medium (-V) at 37°C for 12 days using the CS activity assay kit (Beijing Solarbio Science & Technology Co, Ltd). The sulfide-quinone oxidoreductase (SQR) enzyme was also quantified in strain EGI L10123T using an enzyme-linked immunosorbent assay (ELISA) kit for SQR from Jiangsu Yutong Biotechnology Co, Ltd, during cultivation in 100 mL of chemolithoautotrophic M3 medium (-V) at 37°C for 6 and 12 days.

Complete genome sequencing and analysis of EGI L10123T

Genomic DNA was extracted using the TIANamp Bacteria DNA Kit and sent to Guangdong Magigene Biotechnology Co Ltd for Illumina short-read sequencing as well as Beijing Biomarker Technologies Co, Ltd for PacBio long-read sequencing. The complete genome of strain EGI L10123T was assembled using Unicycler software (ver. 0.5.0) based on a hybrid assembly [42]. Protein-coding regions were predicted using Prodigal (ver. 2.6.3) [34]. Functional annotation of EGI L10123T was also performed by querying predicted CDS against the KEGG database (Release 108.1) using DIAMOND (ver. 0.7.9; E-value <1e−5) [36, 37].

Transcriptome sequencing and analysis of EGI L10123T grown under chemolithoautotrophic conditions

For transcriptome sequencing, EGI L10123T cells were grown in 10 L of M3 (-V) medium with bicarbonate as the sole carbon source, sulfide as the sole electron donor, oxygen (from air) as the electron acceptor, and harvested by filtration (0.22 μm membrane). Filters were transferred immediately to liquid nitrogen and stored at −80°C. Total RNA was extracted using the QIAGEN RNeasy PowerSoil Total RNA Kit (MOBIO). Quality checking, genomic DNA digestion, ribosomal RNA removal, cDNA synthesis, and library construction were performed, and transcriptome data were sequenced by using the NovaSeq 6000 (Illumina) instrument with PE150 at Novogene. Raw transcriptome data were pre-processed as described for the genomic data, and rRNA sequences were removed by using SortMeRNA (ver. 4.3.6) [43]. Finally, the filtered transcriptomic reads were mapped to strain EGI L10123T genes by Salmon (ver. 1.10.1) [44].

Stable isotope evaluation of the carbon fixation pathway in strain EGI L10123T

A 13C-labeling experiment was conducted to detect isotopologues of intermediate metabolites to determine the specific carbon fixation pathway in the chemolithoautotrophic M3 medium (-V) containing 200 g/L NaH13CO3 as the sole carbon source, sulfide (Na2S) as the sole electron donor, and oxygen (from air) as an electron acceptor under oxic conditions (under normal atmospheric pressure conditions). Cultivation was conducted in 30 1000 mL conical flasks (each bottle containing 500 mL medium, in total 15 L) with 150 rpm agitation at its optimal growth temperature of 37°C, using strain EGI L10123T after two consecutive rounds of transfer growth in M3 medium (-V). All conical flasks were rinsed twice with 6 M HCl, followed by three rinses with Milli-Q water, autoclaved, and dried at 60°C before use. The cells were harvested by filtration through a 0.22 μm membrane after 7 days of incubation for isotopologue profiling based on liquid chromatography-mass spectrometry (LC–MS) analysis. The method for LC–MS analysis and calculation of excess 13C were done as described previously [45, 46]. Glutamate, aspartate, and serine were selected for the isotopologue analysis because they are synthesized from the precursors generated by the roTCA [45]. Although alanine is another roTCA indicator, the EGI L10123T strain is missing the alanine aminotransferase (ALT) gene, which means it cannot synthesize alanine from pyruvate and therefore 13C-labeled alanine would not be expected as an intermediate.

Phylogenetic analysis

The complete genome of strain EGI L10123T together with a reference actinobacterial genomic dataset [14] and 167 high-quality Acidimicrobiia genomes (completeness >95% or contamination <5%) from the GTDB database (Release 214.1) were used for phylogenetic analysis to determine the phylogenetic position of strain EGI L10123T. A maximum-likelihood phylogenetic tree was constructed using a multiple sequence alignment of 120 bacterial marker genes generated by GTDB-Tk (ver. 2.3.0) [20] by using IQ-Tree software (ver. 2.2.2.7) [21] with parameters (−alrt 1000 -bb 1000 AUTO). The best-fit model (LG + F + I + I + R10) was well-supported by the Bayesian Information Criterion (BIC).

For phylogenetic analysis of proteins of interest, datasets were derived from the literature and the NCBI prokaryotic gene database. Reference amino acid sequences for ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO) large subunit (RbcL) and NiFe hydrogenases (group1b, 1g, 1h, 2a, 3b, 3d, 4b, 4f) were obtained from previous studies [47–49]. Protein sequences were aligned using MUSCLE (ver. 3.8.31) [50] with 100 iterations, and divergent regions were removed using TrimAL (ver. 1.4) [51]. Phylogenetic inference was performed using IQ-Tree (ver. 2.2.2.7) [22] with parameters (−alrt 1000 -bb 1000 AUTO) and the best-fit model (RbcL: LG + F + R10, [NiFe] hydrogenases groups 1, 2, and 3: LG + R10, as well as [NiFe] hydrogenases group 4: VT + F + R10). The chosen model was well-supported by the BIC. All trees were visualized and annotated using Interaction Tree of Life (iTOL, https://itol.embl.de/).

The evolutionary history of Acidimicrobiia was inferred by using COUNT [52] as previously described [14, 53, 54]. MrBayes (ver. 3.2.0) was used for constructing the Bayesian tree with parameters (ngen = 2 000 000 Nruns = 2 Nchains = 4 diagnfreq = 1000 relburnin = yes burninfrac = 0.25 261 samplefreq = 100 printfreq = 100) using the multiple sequence alignment of 120 bacterial marker genes generated by GTDB-Tk (ver. 2.3.0). The results of the standard deviation of split frequencies (<0.05), the potential scale reduction factor (PSRF = 1), and the effective sample size (ESS > 100) made the Bayesian tree highly reliable. The time-calibrated phylogenetic tree was further inferred by using R (http://cran.r-project.org/, ver. 4.3.1) with the ape package. All trees were visualized and annotated by using iTOL (https://itol.embl.de/).

Distribution and abundance of Acidimicrobiia

Full-length 16S rRNA gene sequences from each Acidimicrobiia genome were submitted to the IMNGS online server (https://www.imngs.org/) [55] for searching against all 16S rRNA gene amplicon datasets from the NCBI Sequence Read Archive (SRA), using a minimum identity threshold of 99%. Samples were removed if the relative abundance of target sequences was <0.1%. Global distribution display was carried out in the R program (http://cran.r-project.org/, ver. 4.3.1).

Results and discussion

Relative abundance of Acidimicrobiia increased after chemolithoautotrophic enrichment

Chemolithoautotrophic microorganisms are believed to contribute significantly to primary production within saline lakes [56]. To study aerobic chemolithoautotrophs in this environment (Fig. S1), oxic enrichments were set up using bicarbonate as the sole carbon source, sulfide as the electron donor, and Barkol Lake sediments as the inoculum (Fig. 1A). Following a 15-day enrichment at 37°C, 16S rRNA gene amplicon (V4 region) analysis revealed changes in the microbial community composition compared to the original sediment samples (Fig. 1B). We found that Acidimicrobiia, which initially had a low relative abundance (day 0), increased following chemolithoautotrophic enrichment (day 15) (Fig. 1C). It has been previously suggested that some Acidimicrobiia possess the potential for carbon fixation via the CBB cycle based on the presence of genes encoding hydrogenase and RuBisCO enzymes (including oxygen-producing photoautotrophic RuBisCO type IA and chemoautotrophic RuBisCO type IE) in several Acidimicrobiia MAGs [14–16, 18, 19], but to our knowledge autotrophy has never been experimentally verified in Acidimicrobiia. However, our analysis of three high-quality Acidimicrobiia MAGs obtained from the sediments failed to uncover key genes (rbcL, rbcS, prk) [57] associated with the CBB cycle, or other intact gene clusters involved in other well-known carbon fixation pathways (Fig. S2 and Table S2). Therefore, we sought to determine whether Acidimicrobiia in general or the specific strains have the capability to be chemolithoautotrophic.

Isolation and identification of a new chemolithoautotrophic Acidimicrobiia family

We performed a culture-dependent strategy [38, 39] to isolate Acidimicrobiia from both the enrichment culture and the original sediment samples. This strategy involved the use of a nutrient medium (0.5 × Marine Agar, 2216E), in combination with a chemolithoautotrophic M3 medium (+V) containing sulfide as the electron donor and bicarbonate as the sole carbon source, both supplemented with 20% agar. A new Acidimicrobiia strain EGI L10123T was isolated from the original sample after 15 days of incubation using a nutrient medium (0.5 × Marine Agar, 2216E) and subsequent repeated colony isolation. Negative staining under transmission electron microscope (TEM) showed that cells of strain EGI L10123T were long rod-shaped (0.4–0.5 μm wide and 5.0–10.0 μm long) (Fig. 2). Moreover, strain EGI L10123T grew at 28–37°C (optimum 37°C), pH 6.0–10.0 (optimum pH 8.0), and in the presence of 0–5% (w/v) NaCl (optimum 0%) using a nutrient medium (Marine Agar, 2216E) (Table S3). These traits are consistent with its environmental source in the saline lake sediment. The complete genome of strain EGI L10123T was obtained and consisted of a single contig 4 107 660 bp bases long, encoding 3931 predicted CDSs, 3 rRNAs (one 23S rRNA, one 16S rRNA, and one 5S rRNA), and 51 tRNAs. The EGI L10123T genome, along with 167 publicly available high-quality Acidimicrobiia MAGs from a variety of environments (Fig. 1D), were analyzed to further assess their diversity, physiology, and evolution. The phylogenetic position and distinctness of EGI L10123T based on relative evolutionary divergence (0.3163809616) indicated that EGI L10123T is a new species of a new family (d__Bacteria; p__Actinobacteriota; c__Acidimicrobiia; o__Acidimicrobiales; f__JACDCH01; g__; s__) within the order Acidimicrobiales (Fig. 1D and Table S4). Here, we propose the name Salinilacustristhrix flava EGI L10123T as the pure culture of the new family Salinilacustritrichaceae. For a comprehensive analysis of the taxonomy and description of strain EGI L10123T, refer to Supplementary Information, Figs S3S7, and Tables S5S6. Although vitamins are typically considered growth factors rather than carbon sources, to exclude the possibility of growth on trace organic compounds in the vitamin solution, cultivation was also conducted in a M3 medium without the addition of trace vitamins. Therefore, chemolithoautotrophic growth of strain EGI L10123T was confirmed by cultivating it in M3 medium with bicarbonate as the sole carbon source and sulfide as the sole electron donor, with (+V) or without trace vitamins (−V), under oxic and anoxic conditions. The results showed that the strain was able to grow in the presence/absence of vitamin solution under oxic conditions (Fig. S8). Subsequently, after three successive transfers, strain EGI L10123T maintained growth in M3 medium (−V) under oxic conditions (atmospheric air, Fig. 2), showing that this new strain is capable of chemolithoautotrophy.

Figure 2.

Figure 2

Verification of chemolithoautotrophy of strain EGI L10123 T . Transmission Electron Micrograph (TEM) cell morphology photos of strain Salinilacustrithrix flava EGI L10123T is shown in the upper right corner of the figure. Cells of strain EGI L0123T were long rod-shaped, each about 5.0–10.0 μm long and 0.4–0.5 μm wide. Error bar represents standard deviation.

Multi-omics insights into the roTCA-based carbon fixation by EGI L10123T

To gain insight into possible carbon fixation pathways and other metabolic features of strain EGI L10123T, we reconstructed major metabolic pathways (Fig. 3, Table S7, and Supplementary Information; Figs S9S14). Previous studies suggested that some members of the class Acidimicrobiia are capable of carbon fixation through the CBB cycle, as determined through metagenomic analysis [15]. However, a comprehensive genome analysis of strain EGI L10123T revealed the absence of key genes that are required for well-known chemolithoautotrophic pathways, including rbcL, rbcS, and prk of the CBB cycle [15, 58]. These genes were also sparse across 168 high-quality Acidimicrobiia genomes, suggesting that the CBB cycle is not a prevalent carbon fixation pathway within this class (Fig. S15). A phylogenetic tree of the 1,5-ribulose bisphosphate carboxylase (RbcL) was constructed to investigate the evolution of this protein in the class Acidimicrobiia (Fig. S16). This revealed a sparse distribution of Acidimicrobiia RbcL homologs across two distinct forms: Form I (A and E), which is responsible for oxygenic photoautotrophs and chemoautotrophs via the CBB cycle, and Form IV RuBisCO-like proteins (RLP), which are not known to catalyze ribulose 1,5-bisphosphate-dependent CO2 fixation [47]. These results implied that the presence of the rbcL gene in this class was due to multiple horizontal gene transfers from distinct microbial taxa. This result suggests that the CBB cycle in this class is not a prevalent carbon fixation pathway. The absence of the CBB cycle in the EGI L10123T genome implies that the other carbon fixation pathways may be functioning under the chemolithoautotrophic condition.

Figure 3.

Figure 3

Overview of metabolic capabilities of strain EGI L10123 T . Different module represents different metabolic modules. Cit: Citrate, Iso: Isocitrate, 2-Oxo: 2-Oxoglutarate, Succ-CoA: Succinyl-CoA, Succ: Succinate, Fum: Fumarate, Mal: Malate, Oxa: Oxaloacetate, PRPP: 5-Phospho-alpha-D-ribose 1-diphosphate, ThPP: Thiamine pyrophosphate, PEP: Phosphoenolpyruvate. This figure was generated according to the gene data in Table S7 using Adobe Illustrator 2022 software.

Despite the absence of key genes for well-known chemolithoautotrophic pathways, we noticed that EGI L10123T encodes CS (cs, gltA) as well as the complete TCA cycle, suggesting that the reversed oxidative TCA (roTCA) may be functioning based on recent research findings that documented the reversibility of CS reactions for carbon fixation [45, 59, 60] (Fig. 3 and Table S7). Transcriptomic analysis of strain EGI L10123T following growth on M3 medium lacking vitamins (−V) with bicarbonate as the sole carbon source and sulfide as the sole electron donor revealed multiple genes associated with the roTCA in the top 30 most abundant transcripts (Tables S8 and S9), including pyruvate ferredoxin oxidoreductase (POR), 2-oxoglutarate ferredoxin oxidoreductase (OGOR), and lower amounts of all other transcripts related to the roTCA (Tables S8 and S9). We also detected the activity of CS (116.6 U/g; U/g is defined as the amount of enzyme that catalyzes the production of 1 nmol of 5-thio-2-nitrobenzoic acid per minute per gram of cells in the reaction system) in strain EGI L10123T grown under chemolithoautotrophic conditions in M3 medium (−V) at 37°C. Together these results implied that strain EGI L10123T may employ the roTCA to fix CO2 [45, 59, 60].

SIP-metabolomics supported involvement of the roTCA in carbon fixation of strain EGI L10123T

To test for carbon fixation via the roTCA in strain EGI L10123T, SIP-metabolomics was carried out using cells grown chemolithoautotrophically with a high concentration (200 g/L) of NaH13CO3 as the sole carbon source and sulfide as the electron donor. Cells were harvested during late-exponential growth phase, and the labeling pattern of proteinogenic amino acids was examined using LC–MS. The number of 13C-labeled carbon atoms in the three amino acids, glutamate, aspartate, and serine were measured as described previously [45], which are direct products of the TCA cycle intermediates 2-oxoglutarate, oxaloacetate, and phosphoenolpyruvate (Fig. 4A and Table S10). The results showed a high amount of 13C-labeled (M + 1) glutamate (10.07%), (M + 2) glutamate (70.72%), (M + 3) aspartate (42.08%), (M + 1) serine (11.67%), (M + 2) serine (64.54%), and (M + 3) serine (20.32%) in NaH13CO3-grown cells. These results provide unequivocal evidence of a functional roTCA in strain EGI L10123T (Fig. 4B and C). The presence of isotopologues indicates ongoing carbon fixation, indicative of the assimilation of 13CO2, whereas heavier isotopologues (≥M + 3) indicate carbon fixation during multiple rounds of the roTCA. The initial reaction of the roTCA is catalyzed by CS, converting citrate into acetyl-CoA and oxaloacetate. Moreover, previous studies have suggested that the detection of aspartate molecules labeled with three or four 13C indicates that oxaloacetate undergoes multiple cycles in the reductive tricarboxylic acid (roTCA) [59]. Although heavy isotopologues (M ≥ 4) were not detected in this study, possibly due to limited detection or interference with other metabolites, the growth of strain EGI L10123T under chemolithoautotrophic conditions, the high expression of key roTCA genes (ppc, por, and kor) (Table S9), and the detection of other key isotopologues together support the possibility of the roTCA being involved in carbon fixation of strain EGI L10123T.

Figure 4.

Figure 4

Analysis of labeling experiments to track the roTCA activity. (A) The incorporation of 13CO2 into serine, aspartate, and glutamate after several rounds of the roTCA produces diverse isotopologues. (B) LC–MS peak area (mAU*min) of the detected isotopologues. (C) The relative abundances of the detected isotopologues. The labeled carbons are schematically represented as circles, in accordance with the legend on the histogram plot. M + X refers to isotopologues carrying X 13C atoms.

Energy metabolism in strain EGI L10123T

Due to the substantial energy and reducing equivalents required by chemolithoautotrophic bacteria for synthesizing organic compounds from inorganic carbon and ATP synthesis, electron donors can become a critical growth-limiting factor [61, 62]. Chemolithoautotrophic prokaryotes rely on inorganic electron donors to provide reducing power for electron transport and ATP synthesis. Given that sulfide was the electron donor in M3 medium (−V), we inferred a role for the SQR, which catalyzes the oxidation of sulfide to sulfite and thiosulfate, in chemolithoautotrophic growth of strain EGI L10123T (Fig. 3). Sulfur quinone oxidoreductase, located on the bacterial cell membrane, catalyzes the oxidation of sulfides into sulfates and sulfite, producing an excess of electrons [63]. These electrons are then transmitted into the electron transport chain, providing ample energy for chemolithoautotrophic carbon fixation [64]. Transcriptomic analysis of strain EGI L10123T under chemolithoautotrophic conditions revealed high expression levels of the sqr gene and electron transport chain genes, including the terminal reductase, cytochrome c oxidase (COX10, COX15, coxA, coxB, and cyoC), and an FoF1-type ATP synthase (Tables S8 and S9). To further probe the expression of SQR, an ELISA was used to quantify SQR proteins in crude extracts from strain EGI L10123T following cultivation in chemolithoautotrophic M3 medium (−V) at 37°C for 6 and 12 days. The results revealed 15 and 25 ng/100 mL, demonstrating the ability of strain EGI L10123T to both transcribe and translate the sqr gene under chemolithoautotrophic conditions (Fig. S17A). We also observed that the concentration of S2− in the culture medium gradually decreased (Fig. S17B). Sulfide (S2−) oxidation by SQR produces various polysulfides (sulfide)n-1, including thiosulfate (S2O32−). The detection results also showed an increase in thiosulfate (S2O32−) concentration, and S2O32− was not detected in the negative control incubations (Fig. S17C). The above evidence further suggested sulfide oxidation, with thiosulfate being one of the primary products resulting from the sulfide oxidation. The transcriptomic analysis of strain EGI L10123T also uncovered high levels of transcripts of key gluconeogenesis genes, including fbp, glpX, pfk, ppdK, and pps (Tables S8 and S9). Carbon fixation transforms carbon atoms from non-carbohydrate precursors into pyruvate or acetone, followed by gluconeogenesis synthesizing two pyruvate molecules into a hexose sugar, which is subsequently metabolized by glycolysis back into acetyl-CoA, providing energy [65, 66]. Together, these results show that chemolithoautotrophy via the roTCA in strain EGI L10123T may be fueled by aerobic sulfide oxidation.

Evolutionary history of chemolithoautotrophy in Acidimicrobiia

With evidence of a sulfide-dependent roTCA in strain EGI L10123T, we sought to better understand the distribution as well as evolution of the roTCA and chemolithotrophy in the Acidimicrobiia. The three key genes for the roTCA, CS, pyruvate:ferredoxin oxidoreductase (POR), and 2-oxoglutarate:ferredoxin oxidoreductase (OGOR) [45], were nearly universal in Acidimicrobiia (Fig. S18). In light of this finding, we speculate that nearly all members of the Acidimicrobiia have the potential to perform CO2 fixation through the roTCA rather than the CBB cycle. Although the thermodynamics of citrate cleavage by CS in the roTCA are unfavorable (free-energy change ∆G′ of more than 35 kJ mol−1) [67], it allows the synthesis of acetyl-CoA from the two molecules of CO2 with reduced ATP expenditure. This characteristic makes the roTCA an energetically efficient chemolithoautotrophic carbon fixation pathway [45, 60]. Additionally, the roTCA pathway has been shown to fix CO2 in a variety of taxa by using on SIP-metabolomics. For example, the roTCA pathway has been demonstrated under high CO2 partial pressures in the acetate-oxidizing and sulfur-reducing Campylobacterota bacterium Desulfurella acetivorans [60], the hydrogen-oxidizing and sulfur-reducing Thermosulfidibacterota bacterium Thermosulfidibacter takaii [68], the Fe (III)-reducing Desulfobacterota species Geobacter sulfurreducens [59], the extremophilic nitrite-oxidizing ChloroflexotaCandidatus Nitrotheca patiens” [69], the Fe (III)-reducing Pseudomonadota bacterium Deferribacter autotrophicus [70], and the marine Campylobacterota species Hippea maritima [45]. The above-mentioned taxa are mainly anaerobic microorganisms, whereas in this study, strain EGI L10123T operates carbon fixation via the roTCA in an exceptionally high bicarbonate medium under oxic conditions. And, we have also detected the highest expression (Transcripts per million, TPM = 5185.286592) of the carbonic anhydrase (Table S8). Carbonic anhydrase dehydrates HCO3 to CO2 provides high levels of CO2 for POR and OGOR for producing oxo-acids. The concentration of CO2 obtained via carbonic anhydrase makes sense in light of the endergonic nature of carbon fixation using these enzymes. This discovery suggests that roTCA may operate both aerobically and anaerobically in environments with high concentrations of bicarbonate or CO2, potentially expanding the habitat range of roTCA functionality.

The utilization of inorganic carbon by chemolithoautotrophic bacteria for the synthesis of organic compounds necessitates a substantial expenditure of energy and reducing equivalents. This process can also help to remediate greenhouse gas emissions by converting inorganic carbon into organic carbon in diverse ecosystems. Inorganic electron donors, such as hydrogen (H2), reduced nitrogen (NH4+ and NO2), and reduced sulfur (e.g. S2, S0, and S2O32−), serve as important sources of reducing power for chemolithoautotrophic metabolism [71]. Hydrogen is recognized as a near-universal electron donor and almost all Acidimicrobiia genomes were predicted to encode hydrogenases (Fig. S19). NiFe hydrogenase groups 3b and 4b were most common in the Acidimicrobiia genomes (Figs S20 and S21). The wide distribution and monophyly of group 4b NiFe hydrogenases from Acidimicrobiia suggest that these genes were likely present in the common ancestor of the class Acidimicrobiia and extended by vertical inheritance. Thus, most Acidimicrobiia may use group 4b NiFe hydrogenases to oxidize formate or carbon monoxide while reducing protons to generate a sodium-motive force through Mrp antiporter modules [48], which is essential for the growth of a variety of halophilic and alkaliphilic bacteria under stressful conditions [72]. Group 3b NiFe hydrogenases are also common and widely distributed in Acidimicrobiia, although they are not monophyletic. These hydrogenases may catalyze NADPH oxidation and the fermentative evolution of H2 [48]. Furthermore, most Acidimicrobiia also encode SQR (Fig. S22), suggesting sulfide as a near-universal electron donor in Acidimicrobiia, although SQR is also known to be used for sulfide detoxification [63, 73]. According to existing research, despite the recognized toxicity of sulfide, it holds metabolic significance as an electron donor for both chemotrophic and photosynthetic organisms [58]. Sulfide quinone oxidoreductase (SQR) oxidizes sulfide into polysulfides and transfers electrons to the electron transport chain of aerobic respiration, generating the proton motive force essential for ATP production [59]. Given the widespread presence of diverse hydrogenases and SQR within the Acidimicrobiia, we suggest that hydrogen and sulfide may be key electron donors for carbon fixation in Acidimicrobiia.

Ancestral character state reconstruction via COUNT was further used to gain insight into the evolution of key roTCA genes, SQR, and diverse hydrogenases across the evolutionary trajectory of Acidimicrobiia (Fig. 5). This analysis indicated that ancestral Acidimicrobiia acquired central metabolic functions related to the roTCA, sulfide oxidation, and hydrogen metabolism before the divergence of five of the orders (Acidimicrobiales, IMCC26256, ATN3, UBA2766, and UBA5794). To test the hypothesis that some of these other Acidimicrobiia can also grow chemolithoautotrophically, we chose three additional Acidimicrobiia strains. These strains were specifically chosen because they lack the CBB cycle and other carbon fixation pathways while possessing all the requisite genes associated with the roTCA (Table S7). These strains were Desertimonas flava DSM 149021T, Actinomarinicola tropica SCSIO 58843T, and Rhabdothermincola sediminis SYSU G02662T. Our experiments demonstrate that these strains can indeed grow under chemolithoautotrophic conditions, regardless of the presence or absence of trace vitamins (using M3 medium with a high concentration (200 g/L) of NaH13CO3 as the sole carbon source and sulfide as the electron donor under normal atmospheric pressure conditions). This is evidenced by a substantial increase in cell numbers observed (Fig. S23). Considering that the key genes for chemolithoautotrophy are ubiquitous in the class Acidimicrobiia, and that other Acidimicrobiia strains lacking the CBB pathway are also capable of chemolithoautotrophy, we speculate that the roTCA, SQR, and diverse hydrogenase constitute integral components of the potential metabolism of chemolithoautotrophy in Acidimicrobiia and may play a pivotal role in its ecological adaptability (Fig. 6A). Our results suggest that Acidimicrobiia thriving in eutrophic environments may have lost their ability to fix CO2 over the long term, leading to an overlook of their potential for chemolithoautotrophic carbon fixation. However, should environmental conditions change drastically, such as through a significant reduction in organic matter, their carbon fixation potential might be retrained and reactivated turning this capability into a crucial one. Further research is needed in the future to determine if these characteristics are actually widespread across the clade.

Figure 5.

Figure 5

Evolutionary history of chemolithoautotrophic function in Acidimicrobiia. The Bayesian tree topology was established using MrBayes [74], employing multiple sequence alignments of 120 bacterial marker genes generated via GTDB-Tk [20]. The Bayesian tree’s robustness was substantiated by the convergence of key indicators: standard deviation of split frequencies (<0.05), PSRF (PSRF = 1), and ESS (ESS > 100). For time-calibrated phylogenetic inference, the R programming language was employed in conjunction with the ape package. The evolutionary trajectory of Acidimirobiia was deduced using COUNT [52], following established methodologies [14, 53], and corroborated via phylogenetic trees constructed for each protein dataset. Star denotes the newly proposed family Salinilacustritrichaceae, within which Salinilacustrithrix flava EGI L10123T is situated.

Figure 6.

Figure 6

Hydrogen- or sulfide-dependent roTCA and distribution of Acidimicrobiia. (A) Model of the hydrogen- or sulfide-dependent roTCA in Acidimicrobiia. (B) Average relative abundance of Acidimicrobiia in different sample types. (C) Global distribution of Acidimicrobiia based on 8349 16S rRNA gene amplicon datasets in which Acidimicrobiia members were present, as a query result from the IMNGS platform [55]. cs: CS gene, por: POR gene, pps: pyruvate, water dikinase gene, ppc: phosphoenolpyruvate carboxylase gene, mdh: malate dehydrogenase gene, fum: fumarate hydratase gene, sdh: succinate dehydrogenase gene, suc: succinyl-CoA synthetase gene, kor/oor: 2-oxoglutarate/2-oxoacid ferredoxin oxidoreductase gene, idh/icd: isocitrate dehydrogenase gene, aco/acn: aconitate hydratase gene.

Globally distribution traits of Acidimicrobiia

The isolation of EGI L10123T and evidence for chemolithoautotrophy described here prompted us to explore the environmental and geographic distribution of chemolithoautotrophic Acidimicrobiia. A total of 94 nearly complete Acidimicrobiia 16S rRNA gene sequences were submitted to the IMNGS platform [55]. In total, 8349 16S rRNA gene amplicon datasets harbored operational taxonomic units with ≥99% similarity to Acidimicrobiia reference sequences, with highest prevalence and relative abundance in aquatic environments as well as sediments (Fig. 6B and C). Considering the evidence of prevalent chemolithoautotrophy in Acidimicrobiia and their widespread distribution in diverse ecosystems, our results imply that Acidimicrobiia could be important, yet underestimated chemolithoautotrophs.

Conclusions

This study provides deep insight into the diversity of Acidimicrobiia as well as the potential of roTCA operating in Acidimicrobiia, driven by sulfide oxidation and probably hydrogen oxidation. Our evidence is based on a combination of genomics, growth experiments, transcriptomics, SIP-metabolomics, and phylogenetics and ancestral state reconstructions. Our study challenges traditional views on carbon fixation within this clade, which previously implied only the CBB based on metagenomes encoding key enzymes of the CBB. Additionally, our discovery may extend the occurrence of the roTCA both phylogenetically and ecologically by showing that the roTCA possibly operates both aerobically and anaerobically in environments with high concentrations of bicarbonate or CO2. This finding improves our understanding of the distribution of the roTCA and its ecological significance.

Protologues

Description of Salinilacustritrichaceae fam. nov. (ICNP)

Salinilacustritrichaceae (Sa.li.ni.la.cus.tri.thri.cha.ce’ae. N.L. fem. n. Salinilacustrithrix, the type genus of the family; −aceae, ending to denote a family; N.L. fem. pl. n. Salinilacustritrichaceae, the Salinilacustrithrix family). The description is the same as for the genus Salinilacustristhrix. The family contains the type genus Salinilacustristhrix.

Description of Salinilacustristhrix gen. nov. (ICNP)

Salinilacustrithrix (Sa.li.ni.la.cus’tri.thrix. N.L. masc. adj. salinus, saline; N.L. masc. adj. lacustris, belonging to a lake; Gr. fem. n. thrix, a hair; N.L. fem. n. Salinilacustrithrix, a hair from a saline lake). The genus description at present is the same as the description of the type species, Salinilacustrithrix flava.

Description of Salinilacustrithrix flava sp. nov. (ICNP)

Salinilacustrithrix flava (fla’va. L. fem. adj. flava, yellow, referring to the color of the colonies).

Cells are long rod-shaped, each about 5.0–10.0 μm long and 0.4–0.5 μm wide. Cells are not motile and have no flagella. Cells are Gram-stain positive and facultatively anaerobic. The major cellular fatty acids are Summed Feature 8 (C17:1  ω6c and/or C17:1  ω7c), iso-C16:0, anteiso-C14:0, and Summed Feature 3 (C16:1ω6c and/or C16:1ω7c). The G + C content of the genomic DNA is 71.81%. Cells contain menaquinones MK-9 (H8). Cells grow at pH values of 6 to 10, temperatures between 28 and 37°C, and NaCl concentrations between 0 and 5%. There is no growth under anoxic conditions without growth factors such as vitamins. The organism is positive for catalase, urease, milk peptonization and coagulation, gelatin liquefaction and coagulation, hydrolysis of cellulose, and degradation of Tweens (20, 40, 60, and 80), and oxidase, but starch negative. The type strain, EGI L10123T (= CGMCC 1.19137T = KCTC 49680T), was isolated from saline lake sediments.

Supplementary Material

Tracked_Supplementary_Information_wrae147
Supplementary_Tables_wrae147

Acknowledgements

We thank Guangdong Magigene Biotechnology Co, Ltd, China, for suggestions of NGS analyses. We also thank Prof Aharon Oren for his recommendation on the nomenclature of the new strain in this study.

Contributor Information

Lei Gao, Key Laboratory of Ecological Safety and Sustainable Development in Arid Lands, Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences, Urumqi, 830011, PR China; State Key Laboratory of Biocontrol, Guangdong Provincial Key Laboratory of Plant Stress Biology and Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), School of Life Sciences, Sun Yat-Sen University, Guangzhou, 510275, PR China.

Lan Liu, State Key Laboratory of Biocontrol, Guangdong Provincial Key Laboratory of Plant Resources and Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), School of Life Sciences, Sun Yat-Sen University, Guangzhou 510275, PR China.

Ai-Ping Lv, State Key Laboratory of Biocontrol, Guangdong Provincial Key Laboratory of Plant Resources and Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), School of Life Sciences, Sun Yat-Sen University, Guangzhou 510275, PR China.

Lin Fu, Key Laboratory of Development and Application of Rural Renewable Energy, Biogas Institute of Ministry of Agriculture and Rural Affairs, Chengdu 610000, PR China.

Zheng-Han Lian, State Key Laboratory of Biocontrol, Guangdong Provincial Key Laboratory of Plant Resources and Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), School of Life Sciences, Sun Yat-Sen University, Guangzhou 510275, PR China.

Takuro Nunoura, Research Center for Bioscience and Nanoscience (CeBN), Japan Agency for Marine-Earth Science and Technology, Yokosuka 237-0061, Japan.

Brian P Hedlund, School of Life Sciences, University of Nevada Las Vegas, Las Vegas, NV 89154, United States; Nevada Institute of Personalized Medicine, University of Nevada Las Vegas, Las Vegas, NV 89154, United States.

Qing-Yu Xu, Key Laboratory of Ecological Safety and Sustainable Development in Arid Lands, Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences, Urumqi, 830011, PR China; State Key Laboratory of Biocontrol, Guangdong Provincial Key Laboratory of Plant Stress Biology and Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), School of Life Sciences, Sun Yat-Sen University, Guangzhou, 510275, PR China.

Dildar Wu, State Key Laboratory of Biocontrol, Guangdong Provincial Key Laboratory of Plant Resources and Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), School of Life Sciences, Sun Yat-Sen University, Guangzhou 510275, PR China.

Jian Yang, Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Wuhan 430074, PR China.

Mukhtiar Ali, Advanced Water Technology Laboratory, National University of Singapore (Suzhou) Research Institute, Suzhou, Jiangsu 215123, PR China.

Meng-Meng Li, State Key Laboratory of Biocontrol, Guangdong Provincial Key Laboratory of Plant Resources and Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), School of Life Sciences, Sun Yat-Sen University, Guangzhou 510275, PR China.

Yong-Hong Liu, Key Laboratory of Ecological Safety and Sustainable Development in Arid Lands, Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences, Urumqi, 830011, PR China.

André Antunes, State Key Laboratory of Lunar and Planetary Sciences, Macau University of Science and Technology, Taipa, Macau SAR 999078, PR China.

Hong-Chen Jiang, Key Laboratory of Ecological Safety and Sustainable Development in Arid Lands, Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences, Urumqi, 830011, PR China; Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Wuhan 430074, PR China.

Lei Cheng, Key Laboratory of Development and Application of Rural Renewable Energy, Biogas Institute of Ministry of Agriculture and Rural Affairs, Chengdu 610000, PR China.

Jian-Yu Jiao, State Key Laboratory of Biocontrol, Guangdong Provincial Key Laboratory of Plant Resources and Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), School of Life Sciences, Sun Yat-Sen University, Guangzhou 510275, PR China.

Wen-Jun Li, Key Laboratory of Ecological Safety and Sustainable Development in Arid Lands, Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences, Urumqi, 830011, PR China; State Key Laboratory of Biocontrol, Guangdong Provincial Key Laboratory of Plant Resources and Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), School of Life Sciences, Sun Yat-Sen University, Guangzhou 510275, PR China.

Bao-Zhu Fang, Key Laboratory of Ecological Safety and Sustainable Development in Arid Lands, Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences, Urumqi, 830011, PR China.

Author contributions

Lan Liu, Jian-Yu Jiao, Bao-Zhu Fang and Wen-Jun Li designed research and project outline. Lei Gao, Bao-Zhu Fang and Yong-Hong Liu obtained the lake sediment sample. Lei Gao, Lan Liu, Ai-Ping Lv, Lin Fu, Qing-Yu Xu, Dildar Wu, Lei Cheng and Meng-Meng Li performed experiments and analysis. Lei Gao, Zheng-Han Lian, Jian Yang and Jian-Yu Jiao performed related bioinformatic analysis. Lei Gao, Bao-Zhu Fang, Jian-Yu Jiao, Takuro Nunoura, Brian P. Hedlund, Lan Liu, Mukhtiar Ali, André Antunes, Hong-Chen Jiang and Wen-Jun Li drafted the manuscript. All authors read and approved the final manuscript.

Conflicts of interest

The authors declare that they have no conflict of interest.

Funding

This research was supported by The Third Xinjiang Scientific Expedition Program (No. 2022xjkk1200), National Science and Technology Fundamental Resources Investigation Program of China (2021FY100900), Natural Science Foundation of Xinjiang Uygur Autonomous Region (2022D01A154), National Natural Science Foundation of China (32000084 and 92251302), the Key-Area Research and Development Program of Guangdong Province (No. 2022B0202110001).

Data availability

The amplicon sequences generated in our study are available in the NCBI SRA database under the BioProject ID PRJNA1012838 with accession number SRR25905149 to SRR25905154. The MAGs described in this paper have been deposited in NCBI database under the BioProject ID PRJNA1012852. The GenBank accession numbers for the 16S rRNA gene sequence and the complete genome of strain EGI L10123T are ON854140 and CP133888, respectively. The transcriptomic sequences of strain EGI L10123T under the chemolithoautotrophic condition are available in the NCBI SRA database under the BioProject ID PRJNA1012866 with accession number SRR25905596.

References

  • 1. Hu  D, Cha  G, Gao  B. A phylogenomic and molecular markers based analysis of the class Acidimicrobiia. Front Microbiol  2018;9:987. 10.3389/fmicb.2018.00987 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Clark  DA, Norris  PR. Acidimicrobium ferrooxidans gen. nov., sp. nov.: mixed-culture ferrous iron oxidation with Sulfobacillus species. Microbiology  1996;142:785–90. 10.1099/00221287-142-4-785 [DOI] [PubMed] [Google Scholar]
  • 3. Johnson  DB, Bacelar-Nicolau  P, Okibe  N  et al.  Ferrimicrobium acidiphilum gen. nov., sp. nov. and Ferrithrix thermotolerans gen. nov., sp. nov.: heterotrophic, iron-oxidizing, extremely acidophilic actinobacteria. Int J Syst Evol Microbiol  2009;59:1082–9. 10.1099/ijs.0.65409-0 [DOI] [PubMed] [Google Scholar]
  • 4. Jones  RM, Johnson  DB. Acidithrix ferrooxidans gen. nov., sp. nov.; a filamentous and obligately heterotrophic, acidophilic member of the Actinobacteria that catalyzes dissimilatory oxido-reduction of iron. Res Microbiol  2015;166:111–20. 10.1016/j.resmic.2015.01.003 [DOI] [PubMed] [Google Scholar]
  • 5. Gavrilov  SN, Korzhenkov  AA, Kublanov  IV  et al.  Microbial communities of polymetallic deposits' acidic ecosystems of continental climatic zone with high temperature contrasts. Front Microbiol  2019;10:1573. 10.3389/fmicb.2019.01573 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. He  YQ, Chen  RW, Li  C  et al.  Actinomarinicola tropica gen. nov. sp. nov., a new marine actinobacterium of the family Iamiaceae, isolated from South China Sea sediment environments. Int J Syst Evol Microbiol  2020;70:3852–8. 10.1099/ijsem.0.004251 [DOI] [PubMed] [Google Scholar]
  • 7. Stackebrandt  E, Rainey  FA, Ward-Rainey  NL. Proposal for a new hierarchic classification system, Actinobacteria classis nov. Int J Syst Evol Microbiol  1997;47:479–91. 10.1099/00207713-47-2-479 [DOI] [Google Scholar]
  • 8. Salam  N, Jiao  JY, Zhang  XT  et al.  Update on the classification of higher ranks in the phylum Actinobacteria. Int J Syst Evol Microbiol  2020;70:1331–55. 10.1099/ijsem.0.003920 [DOI] [PubMed] [Google Scholar]
  • 9. Davis-Belmar  CS, Norris  PR. Ferrous iron and pyrite oxidation by “Acidithiomicrobium” species. Adv Mater Res  2009;71–73:271–4. 10.4028/www.scientific.net/AMR.71-73.271 [DOI] [Google Scholar]
  • 10. Itoh  T, Yamanoi  K, Kudo  T  et al.  Aciditerrimonas ferrireducens gen. nov., sp. nov., an iron-reducing thermoacidophilic actinobacterium isolated from a solfataric field. Int J Syst Evol Microbiol  2011;61:1281–5. 10.1099/ijs.0.023044-0 [DOI] [PubMed] [Google Scholar]
  • 11. Matsumoto  A, Kasai  H, Matsuo  Y  et al.  Ilumatobacter nonamiense sp. nov. and Ilumatobacter coccineum sp. nov., isolated from seashore sand. Int J Syst Evol Microbiol  2013;63:3404–8. 10.1099/ijs.0.047316-0 [DOI] [PubMed] [Google Scholar]
  • 12. Kurahashi  M, Fukunaga  Y, Sakiyama  Y  et al.  Iamia majanohamensis gen. nov., sp. nov., an actinobacterium isolated from sea cucumber Holothuria edulis, and proposal of Iamiaceae fam. nov. Int J Syst Evol Microbiol  2009;59:869–73. 10.1099/ijs.0.005611-0 [DOI] [PubMed] [Google Scholar]
  • 13. Roda-Garcia  JJ, Haro-Moreno  JM, López-Pérez  M. Evolutionary pathways for deep-sea adaptation in marine planktonic Actinobacteriota. Front Microbiol  2023;14:1159270. 10.3389/fmicb.2023.1159270 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Jiao  JY, Fu  L, Hua  ZS  et al.  Insight into the function and evolution of the wood–Ljungdahl pathway in Actinobacteria. ISME J  2021;15:3005–18. 10.1038/s41396-021-00935-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Bay  SK, Waite  DW, Dong  X  et al.  Chemosynthetic and photosynthetic bacteria contribute differentially to primary production across a steep desert aridity gradient. ISME J  2021;15:3339–56. 10.1038/s41396-021-01001-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Norris  PR, Davis-Belmar  CS, Brown  CF  et al.  Autotrophic, sulfur-oxidizing actinobacteria in acidic environments. Extremophiles  2011;15:155–63. 10.1007/s00792-011-0358-3 [DOI] [PubMed] [Google Scholar]
  • 17. Zhao  Y, Liu  P, Rui  J  et al.  Dark carbon fixation and chemolithotrophic microbial community in surface sediments of the cascade reservoirs, Southwest China. Sci Total Environ  2020;698:134316. 10.1016/j.scitotenv.2019.134316 [DOI] [PubMed] [Google Scholar]
  • 18. Hwang  Y, Schulze-Makuch  D, Arens  FL  et al.  Leave no stone unturned: individually adapted xerotolerant Thaumarchaeota sheltered below the boulders of the Atacama Desert hyperarid core. Microbiome  2021;9:234. 10.1186/s40168-021-01177-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Ayala-Muñoz  D, Macalady  JL, Sánchez-España  J  et al.  Microbial carbon, sulfur, iron, and nitrogen cycling linked to the potential remediation of a meromictic acidic pit lake. ISME J  2022;16:2666–79. 10.1038/s41396-022-01320-w [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20. Chaumeil  PA, Mussig  AJ, Hugenholtz  P  et al.  GTDB-Tk v2: memory friendly classification with the genome taxonomy database. Bioinformatics  2022;38:5315–6. 10.1093/bioinformatics/btac672 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Nguyen  LT, Schmidt  HA, Von Haeseler  A  et al.  IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol Biol Evol  2015;32:268–74. 10.1093/molbev/msu300 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Daims  H, Lebedeva  EV, Pjevac  P  et al.  Complete nitrification by Nitrospira bacteria. Nature  2015;528:504–9. 10.1038/nature16461 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23. Jiang  H, Huang  J, Li  L  et al.  Onshore soil microbes and endophytes respond differently to geochemical and mineralogical changes in the Aral Sea. Sci Total Environ  2021;765:142675. 10.1016/j.scitotenv.2020.142675 [DOI] [PubMed] [Google Scholar]
  • 24. Martin  M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J  2011;17:10–2. 10.14806/ej.17.1.200 [DOI] [Google Scholar]
  • 25. Callahan  BJ, Mcmurdie  PJ, Rosen  MJ  et al.  DADA2: high-resolution sample inference from Illumina amplicon data. Nat Methods  2016;13:581–3. 10.1038/nmeth.3869 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Bokulich  NA, Kaehler  BD, Rideout  JR  et al.  Optimizing taxonomic classification of marker-gene amplicon sequences with QIIME 2's q2-feature-classifier plugin. Microbiome  2018;6:90. 10.1186/s40168-018-0470-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Quast  C, Pruesse  E, Yilmaz  P  et al.  The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res  2013;41:D590–6. 10.1093/nar/gks1219 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Bolyen  E, Rideout  JR, Dillon  MR  et al.  Reproducible, interactive, scalable and extensible microbiome data science using QIIME 2. Nat Biotechnol  2019;37:852–7. 10.1038/s41587-019-0209-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Zhang  YM, Wu  G, Jiang  HC  et al.  Abundant and rare microbial biospheres respond differently to environmental and spatial factors in Tibetan hot springs. Front Microbiol  2018;9:2096. 10.3389/fmicb.2018.02096 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Chen  S. Ultrafast one-pass FASTQ data preprocessing, quality control, and deduplication using fastp. iMeta  2023;2:e107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Bankevich  A, Nurk  S, Antipov  D  et al.  SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol  2012;19:455–77. 10.1089/cmb.2012.0021 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Kang  DD, Li  F, Kirton  E  et al.  MetaBAT2: an adaptive binning algorithm for robust and efficient genome reconstruction from metagenome assemblies. PeerJ  2019;7:e7359. 10.7717/peerj.7359 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33. Parks  DH, Imelfort  M, Skennerton  CT  et al.  CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res  2015;25:1043–55. 10.1101/gr.186072.114 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Hyatt  D, Chen  GL, Locascio  PF  et al.  Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinformatics  2010;11:119. 10.1186/1471-2105-11-119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35. Tatusov  RL, Galperin  MY, Natale  DA  et al.  The COG database: a tool for genome-scale analysis of protein functions and evolution. Nucleic Acids Res  2000;28:33–6. 10.1093/nar/28.1.33 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36. Buchfink  B, Xie  C, Huson  DH. Fast and sensitive protein alignment using DIAMOND. Nat Methods  2015;12:59–60. 10.1038/nmeth.3176 [DOI] [PubMed] [Google Scholar]
  • 37. Kanehisa  M, Goto  S. KEGG: Kyoto encyclopedia of genes and genomes. Nucleic Acids Res  2000;28:27–30. 10.1093/nar/28.1.27 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Jiao  JY, Liu  L, Hua  ZS  et al.  Microbial dark matter coming to light: challenges and opportunities. Natl Sci Rev  2020;8:nwaa208. 10.1093/nsr/nwaa280 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Yang  ZW, Lian  ZH, Liu  L  et al.  Cultivation strategies for prokaryotes from extreme environments. iMeta  2023;2:e123. 10.1002/imt2.123 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. Heuer  H, Krsek  M, Baker  P  et al.  Analysis of actinomycete communities by specific amplification of genes encoding 16S rRNA and gel-electrophoretic separation in denaturing gradients. Appl Environ Microbiol  1997;63:3233–41. 10.1128/aem.63.8.3233-3241.1997 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Jiao  JY, Rashidin  A, Zhang  DF  et al.  Advancements in prokaryotic systematics and the role of Bergey's International Society for Microbial Systematics (BISMiS) in addressing challenges in the meta-data era. Natl Sci Rev  2024;11:nwae168. 10.1093/nsr/nwae168 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Wick  RR, Judd  LM, Gorrie  CL  et al.  Unicycler: resolving bacterial genome assemblies from short and long sequencing reads. PLoS Comput Biol  2017;13:e1005595. 10.1371/journal.pcbi.1005595 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43. Kopylova  E, Noé  L, Touzet  H. SortMeRNA: fast and accurate filtering of ribosomal RNAs in metatranscriptomic data. Bioinformatics  2012;28:3211–7. 10.1093/bioinformatics/bts611 [DOI] [PubMed] [Google Scholar]
  • 44. Patro  R, Duggal  G, Love  MI  et al.  Salmon provides fast and bias-aware quantification of transcript expression. Nat Methods  2017;14:417–9. 10.1038/nmeth.4197 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Steffens  L, Pettinato  E, Steiner  TM  et al.  High CO2 levels drive the TCA cycle backwards towards autotrophy. Nature  2021;592:784–8. 10.1038/s41586-021-03456-9 [DOI] [PubMed] [Google Scholar]
  • 46. Steffens  L, Pettinato  E, Steiner  TM  et al.  Tracking the reversed oxidative tricarboxylic acid cycle in bacteria. Bio Protoc  2022;12:e4364. 10.21769/BioProtoc.4364 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. Jaffe  AL, Castelle  CJ, Dupont  CL  et al.  Lateral gene transfer shapes the distribution of RuBisCO among candidate phyla radiation bacteria and DPANN archaea. Mol Bio Evol  2019;36:435–46. 10.1093/molbev/msy234 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Greening  C, Biswas  A, Carere  CR  et al.  Genomic and metagenomic surveys of hydrogenase distribution indicate H2 is a widely utilised energy source for microbial growth and survival. ISME J  2016;10:761–77. 10.1038/ismej.2015.153 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Matheus Carnevali  PB, Schulz  F, Castelle  CJ  et al.  Hydrogen-based metabolism as an ancestral trait in lineages sibling to the Cyanobacteria. Nat Commun  2019;10:463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Edgar  RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res  2004;32:1792–7. 10.1093/nar/gkh340 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Capella-Gutiérrez  S, Silla-Martínez  JM, Gabaldón  T  et al.  trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics  2009;25:1972–3. 10.1093/bioinformatics/btp348 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52. Csurös  M. Count: evolutionary analysis of phylogenetic profiles with parsimony and likelihood. Bioinformatics  2010;26:1910–2. 10.1093/bioinformatics/btq315 [DOI] [PubMed] [Google Scholar]
  • 53. Hua  ZS, Wang  YL, Evans  PN  et al.  Insights into the ecological roles and evolution of methyl-coenzyme M reductase-containing hot spring Archaea. Nat Commun  2019;10:4574. 10.1038/s41467-019-12574-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Jiao  JY, Lian  ZH, Li  MM  et al.  Comparative genomic analysis of Thermus provides insights into the evolutionary history of an incomplete denitrification pathway. mLife  2022;1:198–209. 10.1002/mlf2.12009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55. Lagkouvardos  I, Joseph  D, Kapfhammer  M  et al.  IMNGS: a comprehensive open resource of processed 16S rRNA microbial profiles for ecology and diversity studies. Sci Rep  2016;6:33721. 10.1038/srep33721 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56. Huang  J, Yang  J, Han  M  et al.  Microbial carbon fixation and its influencing factors in saline lake water. Sci Total Environ  2023;877:162922. 10.1016/j.scitotenv.2023.162922 [DOI] [PubMed] [Google Scholar]
  • 57. Banda  DM, Pereira  JH, Liu  AK  et al.  Novel bacterial clade reveals origin of form I Rubisco. Nat Plants  2020;6:1158–66. 10.1038/s41477-020-00762-4 [DOI] [PubMed] [Google Scholar]
  • 58. Garritano  AN, Song  W, Thomas  T. Carbon fixation pathways across the bacterial and archaeal tree of life. PNAS Nexus  2022;1:pgac226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59. Zhang  T, Shi  XC, Ding  R  et al.  The hidden chemolithoautotrophic metabolism of Geobacter sulfurreducens uncovered by adaptation to formate. ISME J  2020;14:2078–89. 10.1038/s41396-020-0673-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60. Mall  A, Sobotta  J, Huber  C  et al.  Reversibility of citrate synthase allows autotrophic growth of a thermophilic bacterium. Science  2018;359:563–7. 10.1126/science.aao2410 [DOI] [PubMed] [Google Scholar]
  • 61. Liang  B, Zhao  Y, Yang  J. Recent advances in developing artificial autotrophic microorganism for reinforcing CO2 fixation. Front Microbiol  2020;11:592631. 10.3389/fmicb.2020.592631 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Srisawat  P, Higuchi-Takeuchi  M, Numata  K. Microbial autotrophic biorefineries: perspectives for biopolymer production. Polym J  2022;54:1139–51. 10.1038/s41428-022-00675-3 [DOI] [Google Scholar]
  • 63. Xia  Y, Lü  C, Hou  N  et al.  Sulfide production and oxidation by heterotrophic bacteria under aerobic conditions. ISME J  2017;11:2754–66. 10.1038/ismej.2017.125 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64. Guzman  MS, Rengasamy  K, Binkley  MM  et al.  Phototrophic extracellular electron uptake is linked to carbon dioxide fixation in the bacterium Rhodopseudomonas palustris. Nat Commun  2019;10:1355. 10.1038/s41467-019-09377-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Köpke  M. Redesigning CO2 fixation. Nat Synth  2022;1:584–5. 10.1038/s44160-022-00131-3 [DOI] [Google Scholar]
  • 66. Wu  C, Lo  J, Urban  C  et al.  Acetyl-CoA synthesis through a bicyclic carbon-fixing pathway in gas-fermenting bacteria. Nat Synth  2022;1:615–25. 10.1038/s44160-022-00095-4 [DOI] [Google Scholar]
  • 67. Guynn  RW, Gelberg  HJ, Veech  RL. Equilibrium constants of the malate dehydrogenase, citrate synthase, citrate lyase, and acetyl coenzyme a hydrolysis reactions under physiological conditions. J Biol Chem  1973;248:6957–65. 10.1016/S0021-9258(19)43346-2 [DOI] [PubMed] [Google Scholar]
  • 68. Nunoura  T, Chikaraishi  Y, Izaki  R  et al.  A primordial and reversible TCA cycle in a facultatively chemolithoautotrophic thermophile. Science  2018;359:559–63. 10.1126/science.aao3407 [DOI] [PubMed] [Google Scholar]
  • 69. Spieck  E, Spohn  M, Wendt  K  et al.  Extremophilic nitrite-oxidizing Chloroflexi from Yellowstone hot springs. ISME J  2020;14:364–79. 10.1038/s41396-019-0530-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70. Slobodkin  A, Slobodkina  G, Allioux  M  et al.  Genomic insights into the carbon and energy metabolism of a thermophilic deep-sea bacterium Deferribacter autotrophicus revealed new metabolic traits in the phylum Deferribacteres. Genes  2019;10:849. 10.3390/genes10110849 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Zhang  SW, Wang  YN, Wang  L  et al.  Role of electron donor in CO2 fixation of chemoautotrophic bacteria and its preconditions: verification in Alcaligenes hydrogenophilus. Enzym Microb Technol  2018;118:37–43. 10.1016/j.enzmictec.2018.06.010 [DOI] [PubMed] [Google Scholar]
  • 72. Lee  Y, Haapanen  O, Altmeyer  A  et al.  Ion transfer mechanisms in Mrp-type antiporters from high resolution cryoEM and molecular dynamics simulations. Nat Commun  2022;13:6091. 10.1038/s41467-022-33640-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73. Marcia  M, Ermler  U, Peng  G  et al.  A new structure-based classification of sulfide:quinone oxidoreductases. Proteins  2010;78:1073–83. 10.1002/prot.22665 [DOI] [PubMed] [Google Scholar]
  • 74. Ronquist  F, Teslenko  M, Van Der Mark  P  et al.  MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol  2012;61:539–42. 10.1093/sysbio/sys029 [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Tracked_Supplementary_Information_wrae147
Supplementary_Tables_wrae147

Data Availability Statement

The amplicon sequences generated in our study are available in the NCBI SRA database under the BioProject ID PRJNA1012838 with accession number SRR25905149 to SRR25905154. The MAGs described in this paper have been deposited in NCBI database under the BioProject ID PRJNA1012852. The GenBank accession numbers for the 16S rRNA gene sequence and the complete genome of strain EGI L10123T are ON854140 and CP133888, respectively. The transcriptomic sequences of strain EGI L10123T under the chemolithoautotrophic condition are available in the NCBI SRA database under the BioProject ID PRJNA1012866 with accession number SRR25905596.


Articles from The ISME Journal are provided here courtesy of Oxford University Press

RESOURCES