BR I E F RE POR T Genome analysis and description of Tunturibacter gen. nov. expands the diversity of Terriglobia in tundra soils Adriana Messyasz1 | Minna K. Männistö2 | Lee J. Kerkhof3 | Max M. Häggblom1 1Department of Biochemistry and Microbiology, Rutgers University, New Brunswick, New Jersey, USA 2Natural Resources Institute Finland, Rovaniemi, Finland 3Department of Marine and Coastal Sciences, Rutgers University, New Brunswick, New Jersey, USA Correspondence Max M. Häggblom, Department of Biochemistry and Microbiology, Rutgers University, 76 Lipman Drive, New Brunswick, NJ 08901, USA. Email: haggblom@rutgers.edu Funding information USDA National Institute of Food and Agriculture, Grant/Award Numbers: Project accession number 1012785, Hatch Project NJ01160; US National Science Foundation, Grant/Award Number: DEB 2129351; Academy of Finland, Grant/Award Numbers: 130507, 310776 Abstract Increased temperatures in Arctic tundra ecosystems are leading to higher microbial respiration rates of soil organic matter, resulting in the release of carbon dioxide and methane. To understand the effects of this microbial activity, it is important to better characterize the diverse microbial communi- ties in Arctic soil. Our goal is to refine our understanding of the phylogenetic diversity of Terriglobia, a common but elusive group within the Acidobacter- iota phylum. This will help us link this diversity to variations in carbon and nitrogen usage patterns. We used long-read Oxford Nanopore MinION sequences in combination with metagenomic short-read sequences to assemble complete Acidobacteriota genomes. This allowed us to build multi-locus phylogenies and annotate pangenome markers to distinguish Acidobacteriota strains from several tundra soil isolates. We identified a phylogenetic cluster containing four new species previously associated with Edaphobacter lichenicola. We conclude that this cluster represents a new genus, which we have named Tunturibacter. We describe four new species: Tunturibacter lichenicola comb. nov., Tunturibacter empetritectus sp. nov., Tunturibacter gelidoferens sp. nov., and Tunturibacter psychrotolerans sp. nov. By uncovering new species and strains within the Terriglobia and improving the accuracy of their phylogenetic placements, we hope to enhance our understanding of this complex phylum and shed light on the mechanisms that shape microbial communities in polar soils. INTRODUCTION Approximately one third of the global soil carbon stock is stored in Arctic soils (Hugelius et al., 2014; Loya & Grogan, 2004), and despite sub-zero temperatures, microbial activity persists throughout the long winters, albeit at a slower rate (Gadkari et al., 2020; Natali et al., 2014; Nikrad et al., 2016; Oechel et al., 1997; Pessi et al., 2022; Poppeliers et al., 2022; Welker et al., 2000). Microbes play important roles in nutrient cycling as they decompose soil organic matter (SOM). As polar soils warm due to climate change, it is of con- cern that enhanced microbial activity will increase the rate of SOM degradation and consequently greenhouse gas emissions (Campbell et al., 2010; Graham et al., 2012; Natali et al., 2014; Welker et al., 2000). SOM degradation usually increases in the short summer season, with cli- mate change resulting in faster and longer soil decompo- sition and increased greenhouse gas emissions that contribute to a positive warming feedback loop (Bond- Lamberty & Thomson, 2010; Koven et al., 2011; Schuur et al., 2008). However, the mechanisms by which climate change affects gas emissions are complex and site- specific, likely determined by for example, vegetation type and microbial community structure (Fry et al., 2023). The microbial community diversity of Arctic environ- ments, like the tundra, is shaped by soil abiotic factors (e.g., pH, SOM, temperature and water activity) and the Received: 3 December 2023 Accepted: 30 April 2024 DOI: 10.1111/1462-2920.16640 ENVIRONMENTAL MICROBIOLOGY This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited. © 2024 The Authors. Environmental Microbiology published by John Wiley & Sons Ltd. Environ Microbiol. 2024;26:e16640. wileyonlinelibrary.com/journal/emi 1 of 19 https://doi.org/10.1111/1462-2920.16640 extreme conditions of the Arctic (e.g., intense UV radia- tion, drought, long periods of sub-zero temperatures interspersed by freeze–thaw events; Männistö et al., 2009; Tas¸ et al., 2018; Tveit et al., 2014; Viitamäki et al., 2022; Weintraub & Schimel, 2003). Despite these harsh conditions, there is tremendous bacterial diversity in tundra soils, with the bacterial communities dominated by Acidobacteriota, Actinomy- cetota, and Pseudomonadota (Proteobacteria; Chu et al., 2010; Kim et al., 2014; Männistö et al., 2007, 2013; Viitamäki et al., 2022; Vorˇísˇkova et al., 2019). Members of these phyla change in abundance, compo- sition, and activity in response to environmental vari- ables such as seasonal changes (freeze–thaw cycles), vegetation, and soil pH (Männistö et al., 2009, 2013; McMahon et al., 2011; Viitamäki et al., 2022; Zinger et al., 2009). The ubiquitous but elusive Acidobacteriota are particularly susceptible to changes in soil pH and different members of the phylum thrive at different pH levels. Members of the Class Terriglobia (formerly Acid- obacteriia; Sub-division 1 [SD1] Acidobacteriota) are particularly abundant in the cold, nutrient-poor, acidic soils such as those in Arctic tundra and boreal forest biomes (Belova, Ravin, et al., 2018; Campbell et al., 2010; Ivanova et al., 2020; Kim et al., 2014; Männistö et al., 2007, 2009, 2018). The Acidobacteriota is one of the most abundant phyla in soils around the globe; however, the diversity within this phylum is still largely uncharacterized (Zhang et al., 2022). Initial culture-independent methods initially delineated 26 subdivisions of Acidobacteria (Barns et al., 2007). More recently, these subdivisions were refined into 15 class-level units (Dedysh & Yilmaz, 2018). Of these, the classes Terriglobia, Blasto- catellia, and Vicinamibacteria are most abundant in soils (Janssen, 2006; Jones et al., 2009). Furthermore, mem- bers of the Terriglobia have been the most successfully cultivated group with over 40 named and published spe- cies to date (Göker, 2023; Zhang et al., 2022). The family Acidobacteriaceae currently encom- passes over a dozen named genera; however with fas- ter and more advanced sequencing technologies a greater number of Acidobacteriota species and strains are being described (Kalam et al., 2020; Zhang et al., 2022). The analysis of multiple genes from isolate genomes and environmental metagenomes has contrib- uted towards a better understanding of the diversity of Acidobacteriota in multiple ecosystems without the need for cultivation (Dedysh et al., 2022; Kalam et al., 2020; Kielak et al., 2016; Parsley et al., 2011). Genome analy- sis has indicated a variety of metabolic functions within the Acidobacteriota including complex carbohydrate degradation (xylan, cellulose, hemicelluloses, pectin, starch, and chitin), nitrogen metabolism (nitrate, nitrite, and nitric oxide reduction), genes for oxygen utilization in hypo- and hyperoxic conditions, as well as the ability to oxidize methanol (Diamond et al., 2019; Eichorst et al., 2007; Kalam et al., 2020; Rawat et al., 2012; Ward et al., 2009). Some Acidobacteriota is equipped with reg- ulatory genes in response to stressors like starvation, oxidative stress, heat stress, as well as acid resistance systems (Challacombe et al., 2011; Kalam et al., 2020; Kielak et al., 2016; Ward et al., 2009). Secondary metabolites from biosynthetic gene clusters were also found within Acidobacteriota. Specifically, non-ribosomal peptide synthetases and polyketide synthases were found within Candidatus Angelobacter (a proposed Ter- riglobia), indicating mechanisms for interbacterial com- petition (Crits-Christoph et al., 2022). There may be several reasons for such a high taxo- nomic and functional diversity of Acidobacteriota in terres- trial systems. Primarily, soils create complex and spatially/temporally heterogeneous environments, with this disequilibrium ensuring that no one species domi- nates over others (Ghoul & Mitri, 2016; Lennon et al., 2012). Additionally, ecological drivers, such as resource partitioning, selective predation, and temporal separation, most likely shape the diversity of Acidobacter- iota and other soil phyla (Chesson, 2000; Chesson & Huntly, 1997). To better understand how Acidobacteriota activity is shaped by environmental and ecological factors we need to identify the yet uncharacterized species/ strains within this phylum and compare their genomes to elucidate metabolic potential. In this study, we sequenced several new tundra soil Acidobacteriota isolates from Malla Nature Reserve, Kilpisjärvi, Finland using both Illu- mina short-read and Oxford Nanopore long-read technol- ogies. Through rRNA operon and full genome analysis, we expand the current phylogenetic diversity of the order Terriglobales in the Acidobacteriota. We also delineate and name a novel genus, Tunturibacter, containing four new species in phylogenetic proximity to other species within the genera Terriglobus (Eichorst et al., 2007; Männistö et al., 2011; Podar et al., 2019; Rawat et al., 2012), Granulicella (Männistö et al., 2012; Pankratov & Dedysh, 2010; Rawat et al., 2013, 2014), and Edaphobacter (Koch et al., 2008; Wang et al., 2016). With functional annotation and comparison of this new genus to known Acidobacteriota, the core and unique genes of the genera within this group can be identified. Our phylogenetic analysis demonstrates that a complex diversity of the Terriglobales can be uncovered. Addition- ally, understanding how novel genera differ within Acido- bacteriota can help further elucidate why such diversity exists in soil environments. EXPERIMENTAL PROCEDURES Acidobacteriota isolation and DNA extraction Acidobacteriota strains were isolated from tundra soil samples collected from Malla Nature Reserve, 2 of 19 MESSYASZ ET AL.ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Kilpisjärvi, Finland (69010 N, 20500 E) in July 2010 (strain codes that begin with ‘M’) and July 2012 (strain codes that begin with ‘X’). Several carbon substrates were tested in an attempt to isolate novel members of the Acidobacteriota. Soil samples were diluted in VL55 mineral medium (Davis et al., 2005) and dilution plated on the different media. Plates were incubated at 4C for up to 3 months and inspected every 2 weeks. To isolate Acidobacteriota and other slow-growing oli- gotrophs, only colonies that formed after 1–2 months of incubation were picked and streaked to new media. After purification, isolated strains were identified by Sanger sequencing of the 16S rRNA gene. DNA was extracted from the isolates using the DNeasy Ultra- Clean Microbial Kit (Qiagen) according to the manufac- turer’s instructions. 16S rRNA genes were amplified using 27f and 1525r primers (Lane, 1991) and the poly- merase chain reaction (PCR) products were sequenced by LGC Genomics (Berlin, Germany) using the forward primer 27 f. Approximately 800 bp sequences were compared with those in the EzTaxon database (Yoon et al., 2017) and those identified as members of the Acidobacteriota were selected for fur- ther studies. Strains M8UP20, M8UP22, M8UP23, M8UP27, M8UP28, M8UP30, and M8UP39 were isolated using a mixture of carboxy methyl cellulose (CMC), xylan, pec- tin, and starch (each 0.25 g L1) in VL55 mineral salt medium amended with yeast extract (0.1 g L1) and agar (20 g L1). Strain MP8S11 was isolated using cel- lobiose, trehalose, and sucrose (each 0.25 g L1) in VL55 mineral salt medium amended with yeast extract (0.1 g L1) and agar (20 g L1). Strains X4BP1, X5P3, X5P6, and X4EP2 were isolated using xylan (0.5 gL1), cellobiose (0.25 g L1), and xylose (0.25 g L1) in VL55 mineral salt medium amended with MEM essen- tial and non-essential amino acids (Sigma-Aldrich®). VL55 mineral salt solution was prepared as described by Davis et al. (2005), pH was adjusted either to 4.0 or 5.5. All strains were maintained at pH 5.5 using GY medium containing glucose (1 g L1) and yeast extract (0.5 g L1) in VL55. Isolation of Granulicella arctica MP5ACTX2 has been described earlier (Männistö et al., 2012). Phenotypic and chemotaxonomic analyses Growth of strains M8UP23, M8UP39, and X5P6 on GY medium was tested at different temperatures and pH. The temperature range was evaluated by growth on GY-agar incubated at 2, 4, 10, 15, 20, 25, and 30C. The plates were checked every 1–2 days for growth. Visual differences in growth at different temperatures were recorded. The effect of pH on growth was evalu- ated by growing the strains in liquid GY medium at pH 3.0–9.0 (in 0.5 pH unit increments) in 96 well plates. Carbon source utilization of the three strains was ana- lysed in 96-well plates for up to 10 days at 20C with VL55 mineral medium supplemented with 100 mg L1 yeast extract and 10 mM of each carbon source. Yeast extract was required for good growth on single carbon sources. The control contained only yeast extract. Growth was measured at 620 nm using a Multiscan FC microplate reader (Thermo Scientific). Hydrolysis of dif- ferent polysaccharides (starch, CMC, xylan, lichenan, pectin, xanthan, and gum arabic) was determined at room temperature by observing CO2 production for up to 20 days and analysed as described in Männistö et al. (2012). Cellular fatty acids were analysed by gas chroma- tography as described in Männistö et al. (2012). Analy- sis of respiratory quinones and polar lipids was carried out by the Identification Service, DSMZ, Braunschweig, Germany. For the analysis, cells were cultivated on GY medium, collected by centrifugation, washed with PBS buffer, and freeze-dried. Polar lipids were extracted from 200 mg of freeze-dried cells and separated by thin-layer chromatography (German Collection of Microorganisms and Cell Cultures GmbH: Polar Lipids, n.d. https://www.dsmz.de/services/microorganisms/bio chemical-analysis/polar-lipids). Respiratory quinones were extracted from 50 mg of freeze-dried cells and analysed by HPLC (German Collection of Microorgan- isms and Cell Cultures GmbH: Respiratory Quinones, n.d. https://www.dsmz.de/services/microorganisms/bioch emical-analysis/respiratory-quinones). Genome sequencing and assembly A total of 13 Acidobacteriota isolates were used for whole genome sequencing via the Oxford Nanopore MinION. Sequence libraries were prepared using R 9.4 chemistry and the MinION Rapid Sequencing kit (SQK-RAD004). Libraries were run on a MinION R9.4 Flongle for 24 h. Fast5 files were basecalled using Guppy (V 6.0.1; Oxford Nanopore Technologies Ltd.) in high accuracy mode and the FastA reads were used for genome assembly with the Trycycler assembler pack- age (V 0.5.3; Wick et al., 2021), which allows for multi- ple assemblies specific to Nanopore long read sequences. For these genomes, the assemblers Flye— nano-hq (V 2.9; Kolmogorov et al., 2019; Lin et al., 2016), Minipolish (V 0.1.2; Wick & Holt, 2019), and Raven (V 1.8.1; Vaser & Sˇikic, 2021) were employed using default settings. The contigs were put into clusters and multiple sequence alignments were generated to produce a consensus or final assembly using Medaka (V 1.6.0; © 2018 Oxford Nanopore Tech- nologies Ltd.). In addition, several of the assemblies were run through PolyPolish (V 0.5.0; Wick & Holt, 2022) to close and complete the genomes. Of the 13 isolates used for TUNTURIBACTER GEN. NOV. IN THE ACIDOBACTERIOTA PHYLUM 3 of 19ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License genome assembly, 6 also had Illumina short-read sequences publicly available via the JGI Genome Por- tal (see Project IDs in Table S1), and 4 (MP8S11, M8UP39, X5P6, and M8UP23) were previously sequenced via Illumina (see Supporting information S1, Methods and Results). These short reads along with our Trycycler-Medaka assemblies were used in the PolyPolish pipeline (default parameters). Genome assembly completeness and contamination for all 13 isolates were analysed via CheckM (V 1.0.18; Parks et al., 2015). Percent GC was calculated via QUAST (V 4.4; Gurevich et al., 2013). For each assembly with greater than one contig, the second longest contig was searched against the plasmid database PLSDB (V 2023_11_03_v2; Schmartz et al., 2021). This search was conducted via the PLSDB online API tool using Mash (V 2.3; Ondov et al., 2016) to search against the PLSDB database (parameters used: maximal p-value 0.1, maximal distance 0.1, minimal identity 0.70). Gene prediction was run via Prokka (V 1.14.5; Seeman, 2014). To find annotations to carbohydrate- active enzymes (CAZy), genes annotated via Prokka were scanned using a set of Hidden Markov Models (HMMs) from the dbCAN2 CAZy collection (dbCAN HMM database v10; Zhang et al., 2018) using HMMER (v 3.3.2; Eddy, 2011). The minimum e-value for this search was 1e15 and the minimum coverage of the model length was set at 90%. The Genome Taxonomy Database (GTDB; releases 207 and 214; GTDB R07-RS207 and R08–RS214; Parks et al., 2022) was used to taxonomically classify the genomes via the GTDB-Tk tool (v 2.3.2). Genome phylogenies and pangenome analysis A single-copy gene (SCG) phylogeny including the 13 new tundra soil isolate genomes and known Class Terriglobia (SD1 Acidobacteriota) species was ana- lysed via the GToTree program (V 1.7.05; Lee, 2019). The program was run with default settings for alignment (MUSCLE V 5.1; Edgar, 2021) and approximately maximum-likelihood phylogenetic tree building (FastTree V 2.1.11; Price et al., 2010). The ‘-H Bacte- ria’ parameter was used to identify 74 bacterial SCGs across all genomes analysed (listed in Supporting information S1 and S2). Another multi-locus tree for the same genomes was built via OrthoFinder (V 2.5.4; Emms & Kelly, 2019). The Orthofinder tree also utilized a MUSCLE (V 5.1; Edgar, 2021) alignment and phylo- genetic tree building with FastTree (Price et al., 2010). Both SCG and OrthoFinder trees were viewed via the iTOL interface (V6; Letunic & Bork, 2021). These trees were compared with 16S rRNA gene and rRNA-operon trees, which included a more complete representation of the Terriglobia and the novel Acidobacteriota isolates included in the rRNA operon analysis (see Supporting information S1, Methods and Results). Genome align- ment visualizations were conducted via PGV-Mummer (default setting; V 0.3.2; Shimoyama, 2022) and pro- gressive MAUVE (Geneious plugin V 1.1.1) with 15 match seed weight, minimum 5000 Locally Collinear Blocks score, full alignment, and MUSCLE 3.6 gapped aligner (Darling et al., 2004). Genome assemblies of the new tundra isolates along with some select Terriglobia were further com- pared through the Anvi’o pangenome analysis pipeline (Eren et al., 2021). A contig database for each genome was created with the following additions: HMM search, SCG taxonomy, tRNAs scan, NCBI cogs search, and KEGG Kofam search. Average nucleotide identity across the compared genomes was calculated via pyANI (Pritchard et al., 2015) program plugin within the Anvi’o pangenome pipeline (command: ‘anvi-compute-genome-similarity’). The pangenome was then visualized via the Anvi’o interactive display, organizing samples by a gene cluster frequencies tree. Functional enrichment between strain categories within the pangenome analysis was computed via anvi-com- pute-functional-enrichment-in-pan using the KEGG_- Class and COG20_PATHWAY annotation sources (Shaiber et al., 2020). Analysis of Acidobacteriota community in Malla tundra heath soils Soil samples were collected from tundra heaths of Mt. Pikku Malla in Malla Nature Reserve, Kilpisjärvi (69030 5000N, 204404000 E), with differences in topogra- phy that dramatically influence snow accumulation. Four plots representing windswept slopes and four plots corresponding to snow-accumulating biotopes were sampled at a depth of <5 cm in February 2013 as described previously (Männistö et al., 2024). Compos- ite soil samples of five soil cores were taken from each plot with three subsamples from each composite sam- ple used for DNA extraction with a CTAB-based method. Near full-length bacterial rRNA operons were ampli- fied from extracted DNA using 16S rRNA-27 Forward and 23S rRNA-2241R primers, <10 ng template DNA, and a High-Fidelity Taq Polymerase (Biomake Inc., CA, USA; Kerkhof et al., 2017). PCR conditions were: Initial denaturation for 3 min at 94C; followed by 25 cycles of 98C/10 s denaturation, 60C/15 s primer annealing and 72C/240 s extension; and a final extension at 72C for 5 min. Barcoded rRNA operon amplicons were visualized and quantified by agarose gel electrophore- sis and stored at 20C until library preparation. Library construction utilized the SQK-LSK108 sequencing kit and sequencing via the Oxford Nanopore MinION. The fast5 files were basecalled using Guppy (3.2.0). Raw 4 of 19 MESSYASZ ET AL.ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License reads were demultiplexed with Guppy and sized (3700–5000 bp) using Geneious (11.1.5). FastA files were initially screened via MegaBLAST (2.10.0) against the ribosomal RNA operon database (rOPDB; Kerkhof et al., 2022) to determine the raw reads associated with the Acidobacteriota. These Acidobacteriota reads were re-screened against a modified database using the original Acidobacteriota rRNA operons from the rOPDB, amended with rRNA operons from the new Tunturibacter and Granulicella strains described in this study. Best BLAST hits (BBHs) were identified using the following settings: word size 60, match/mismatch cost 2/3, gap open/extend 0/4, and e-value of 1  1010. Relative abundance of the Acidobacteriota BBHs to the updated database was calculated and bubble plots were generated in RStudio (2023.06.0 + 421) using ggplot2 and reshape2. RESULTS AND DISCUSSION Novel tundra Acidobacteriota isolates Acidobacteriota strains were isolated from tundra heath soil samples from Malla Nature Reserve using different carbon substrate combinations. Initial analysis of partial 16S rRNA sequences (data not shown) indicated that these were members of the Terriglobia, closely related to Edaphobacter and Granulicella species. Physiologi- cal comparisons of our isolates to known Edaphobacter and Granulicella strains showed similarities in the utili- zation of various organic compounds and slight differ- ences in temperature and pH range. Further investigation utilizing long-read Oxford Nanopore sequencing to assemble full genomes uncovered phy- logenetic differences between the isolates and other members of the Terriglobia. The consistent phyloge- netic placement across several comparative tools (sin- gle-gene trees, multi-locus trees, pangenome analysis) clearly separates the tundra heath isolates from known Edaphobacter and Granulicella species. Therefore, these isolates ultimately represent a novel genus within Terriglobia for which we propose the name Tunturibac- ter. Here, we present the genomic, phylogenetic, and phenotypic characterization of four novel species within the Tunturibacter genus. Genome assemblies The 13 new Acidobacteriota genome assemblies are listed in Table 1. Of these, 10 included Illumina short reads used to polish the genomes to >99% completion. Three of the 10 complete genomes had only one scaf- fold/contig, while the remainder had two or more scaf- folds/contigs, with the longer scaffold representing the bacterial chromosome and the shorter contigs representing smaller genetic elements, that is, plas- mids. To examine whether these shorter contigs could be plasmids, we searched the shorter contigs from assemblies with more than one contig against the PLSDB plasmid database. For each assembly, the sec- ond longest contig had one or more hits with an aver- age of 75% identity to an already annotated plasmid from the database (Table S2). Their predominant align- ment to plasmids from Granulicella tundricola at only around 75% identity indicates these are novel plasmids belonging to the Tunturibacter. These plasmids may code for a few extra genes that are not necessary for Tunturibacter growth, reproduction, or existence, but may offer some resistance strategies beneficial for survival. The assembled genomes of the Tunturibacter strains had between 3600 and 7400 predicted genes, of which >98% were protein-coding genes for each assembly (Table 1). Of these genes, on average 44% were annotated as non-hypothetical. Full assembly sta- tistics for these genomes are included in Table S1. Based on GTDB annotation, 10 out of 13 assemblies were classified as most similar to Edaphobacter licheni- cola DSM 104462 with ANI values ranging from 89% to 100% and amino acid (AA) similarities from 76% to 94%. Of the remaining three assemblies, two were classified as most similar to G. arctica MP5ACTX2 (DSM 23128; 79%–100% ANI and 86%–94% AA) and one to Granulicella mallensis MP5ACTX8 (100% ANI and 94% AA). Multi-locus phylogeny of the SD1 Acidobacteriota To determine the taxonomic placement of the new tun- dra Acidobacteriota isolates we compared 16S rRNA gene, rRNA operon, and single copy gene phylogenies. Utilizing the full genomes of members of the Terriglo- bia, we phylogenetically compared the genome assem- blies using 74 selected SCGs (Figure 1A). The resulting SCG tree more accurately groups the Eda- phobacter, Granulicella, and Terriglobus species than the rRNA operon or 16S rRNA gene trees alone (Figures S1A,B and S2A,B). Importantly, the assem- blies grouping together with E. lichenicola DSM 104462 were phylogenetically separated from the other Eda- phobacter species, including the type species Edapho- bacter modestus, and would thus represent the new genus Tunturibacter. As can be seen from the SCG tree, this new genus appears to comprise four species, each with several strains. The phylogeny is more evident in the SCG sub-tree showing the members of the Terriglobia most closely related to the proposed Tunturibacter genus (Figure 1B). The SCG tree also more accurately clus- ters the Granulicella species. Another potential new TUNTURIBACTER GEN. NOV. IN THE ACIDOBACTERIOTA PHYLUM 5 of 19ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License T A B L E 1 G en om e as se m bl ie s of th e 13 A ci do ba ct er io ta is ol at es fr om M al la N at ur e R es er ve ,K ilp is jä rv i, F in la nd . C h ar ac te ri st ic s S tr ai n M 8U P 23 M 8U P 22 M 8U P 27 M 8U P 20 M 8U P 39 M 8U P 30 M 8U P 28 X 5P 6 X 4B P 1 M P 8S 11 X 5P 3 X 4E P 2 M P 5A C T X 2 % C om pl et io n 99 .7 8 99 .7 8 10 0 89 .3 1 10 0 10 0 91 .4 2 10 0 10 0 10 0 10 0 10 0 95 .5 2 C on ta m in at io n 0 0 0. 05 0 0. 86 0. 86 0. 86 0. 86 0. 86 1. 72 2. 63 1. 71 0. 86 % G C 57 .8 2 57 .7 6 58 .4 3 58 .4 1 57 .4 2 57 .5 6 57 .0 8 56 .7 2 56 .7 56 .2 6 57 .8 3 58 .2 6 57 .6 8 T ot al nu m be r of ge ne s pr ed ic te d 42 80 43 77 41 99 53 45 48 35 45 08 74 22 48 45 53 86 53 67 52 66 36 96 56 46 % P ro te in co di ng ge ne s 98 .7 4 98 .7 9 98 .6 7 98 .9 3 98 .8 6 98 .8 2 99 .2 9 98 .9 3 98 .9 6 98 .9 9 99 .0 1 98 .5 9 99 .0 8 % G en es w ith no n- hy po th et ic al fu nc tio n 44 .3 9 44 .1 6 44 .6 1 38 .0 2 41 .8 4 43 .0 6 35 .2 9 42 .8 4 41 .5 7 41 .3 3 43 .3 5 47 .3 5 38 .2 4 T ot al nu m be r of C A Z y ge ne s 48 6 51 3 51 9 47 1 54 2 50 4 61 1 68 7 70 3 63 7 83 4 44 4 56 1 N um be r of co nt ig s 3 2 2 2 2 2 2 2 1 2 1 1 6 C hr om os om e to ta l si ze (b p) 4, 89 5, 84 9 5, 07 3, 71 7 4, 84 1, 14 9 4, 83 0, 04 5 5, 45 0, 51 8 5, 14 8, 40 8 6, 61 6, 84 3 5, 61 6, 89 1 6, 12 9, 46 0 5, 51 7, 69 3 6, 50 9, 45 9 4, 30 5, 84 7 5, 27 2, 91 8 S ec on d lo ng es t co nt ig si ze (b p) 11 4, 87 4 15 ,0 20 55 ,3 09 78 ,1 40 12 8, 51 9 13 8, 88 3 21 3, 90 6 16 ,6 51 N a 75 6, 04 6 N a N a 10 1, 20 9 6 of 19 MESSYASZ ET AL.ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License F I GURE 1 (A) Single copy gene phylogeny of Terriglobia genomes. (B) Sub-tree of the phylogenetic placement of the new Tunturibacter genus with more closely related genera. Accession numbers for genomes available from NCBI are listed in Table S3. TUNTURIBACTER GEN. NOV. IN THE ACIDOBACTERIOTA PHYLUM 7 of 19ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License genus may be represented by G. mallensis, Granuli- cella paludicola, and Granulicella cerasi, which group with Bryocella elongata, phylogenetically distinct from the other species in the Granulicella cluster, including the type species G. paludicola (Figure 1A). A compari- son of these genomes via a multi-locus phylogeny based on orthogroups and orthologs (Orthofinder) also resulted in a similar phylogenetic structure as the SCG tree (Figure S3). Overall, compared with the rRNA operon tree, phy- logenies built on multiple genes from fully assembled genomes more accurately delineated the Acidobacter- iota genera and species. However, the rRNA operon tree was still able to differentiate the various novel clus- ters, revealing the polyphyletic nature of some of these genera, supporting the separation of E. lichenicola and three new species as the proposed new genus Tunturibacter. Pangenome analysis Pangenome analysis was used to compare gene fre- quencies from the COG, KEGG, SCG, and Kofam annotations of the input genomes (Figures 2 and 3). The gene frequency tree created from this analysis (Figure 2, right side) is similar to the SCG tree, again supporting the separation of two new potential genera in the Terriglobia. The proposed Tunturibacter genus is present as its own cluster, as is the cluster which groups several Granulicella species that affiliate with B. elongata (Figure 2). The very last row indicating the F I GURE 2 Pangenome analysis of the Terriglobia in the Acidobacteriota. Comparison of gene clusters between the 13 assembled tundra soil isolates and known Terriglobia. Each row designating an Acidobacteriota strain starts with information on the presence of gene clusters (gene clusters are marked by the darker-coloured regions within the row). This is followed by the red/maroon columns indicating levels of genome total length, number of genes per kbp, singleton gene clusters, and number of gene clusters. These columns are followed by the ANI data from the pangenome analysis, and the grouping of the Acidobacteriota based on gene cluster frequency (right-hand tree). Under the gene cluster presence rows are rows indicating COG20, KEGG, and Kofam annotations, SCG clusters, and the number of genes and contributing genomes in the gene cluster. 8 of 19 MESSYASZ ET AL.ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License number of contributing genomes (Figure 2, bottom left) shows that the area of the pangenome comparison with the most contributing genomes also includes the SCG clusters, representing the core genome of the Acido- bacteriota analysed (Figure 3; Bin 1, 337 clusters, 11,791 gene calls). Except for strain M8UP15, ANI analysis clearly sep- arated the proposed new Tunturibacter species from the other Edaphobacter species, with E. modestus as the type species. The ANI similarities, along with the phylogenetic clusters of this new genus in the SCG tree (Figure 1) support the separation of four species within the new Tunturibacter genus with their respective type strains: E. lichenicola DSM 104462 transferred to the genus as Tunturibacter lichenicola comb. nov. (includes strain MP8S11), Tunturibacter psychrotoler- ans sp. nov. X5P6 (includes strains X4BP1 and X5P2), Tunturibacter empetritectus sp. nov. M8UP23 (includes strains M8UP22, M8US30, M8UP20, and M8UP27), and Tunturibacter gelidoferens sp. nov. M8UP39 (includes strains M8UP30 and M8UP28). Bin 2 captures the gene clusters unique to this new genus (Figure 3; Bin 2, 147 clusters, 1903 gene calls). One major difference in this analysis compared with the SCG tree, is the grouping of genome M8UP15 with E. aggregans DSM 19364 rather than the Tunturi- bacter genus. These two genomes have the highest number of gene clusters, with one area sharing F I GURE 3 Circularized Anvio pangenome of the Terriglobia. TUNTURIBACTER GEN. NOV. IN THE ACIDOBACTERIOTA PHYLUM 9 of 19ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License multiple genes unique to these two genomes (Figure 3; Bin 3, 1820 clusters, 3640 gene calls). The high ANI similarity of these two genomes indicates that strain M8UP15 is most closely related to E. aggregans DSM 19364. However, out of all the genomes in this analysis, M8UP15 and E. aggregans DSM 19364 share the highest number of partial genes (416), indi- cating that their genomes are not complete. Since the SCG and operon trees did not group these two genomes closely, it is unclear whether this result is driven by their unique genes or their incomplete genomes. The taxonomic placement of strain M8UP15 may thus require future reassessment. Physiological comparison of Tunturibacter species The phenotypic properties of the proposed type-strains for the new species, Tunturibacter empetritectus M8UP23, T. gelidoferens M8UP39, and T. psychroto- lerans X5P6 were compared with Tunturibacter (Eda- phobacter) lichenicola (Belova, Suzina, et al., 2018). Similarities were found for all strains based on their utili- zation of various organic compounds (Table 2), as well as hydrolysis of polysaccharides, enzyme activities (API ZYM tests), and composition of quinones and major cellular fatty acids (Table S4). The temperature growth range for strains T. empetritectus M8UP23, T. gelidoferens M8UP39, and T. psychrotolerans X5P6 was 2–30C, while the growth range of T. lichenicola DSM 104462 was different at 7–37C. These Tunturi- bacter species/strains tolerate cold conditions, similar to Terriglobus (4–30C; Eichorst et al., 2007; Männistö et al., 2011) and Granulicella species (4–28C; Männistö et al., 2012). The growth pH range for T. empetritectus M8UP23, T. gelidoferens M8UP39, and T. psychrotolerans X5P6 was 3.5–6.5 and for T. licheni- cola DSM 104462 was 3.4–7.0. These ranges are most similar to those of G. tundricola and G. mallensis (pH 3.5–6.5; Männistö et al., 2011) and can tolerate more acidic conditions than Terriglobus (pH 4.5–7.5; Eichorst et al., 2007; Männistö et al., 2012). Genome comparisons of Tunturibacter species Genome alignments at various taxonomic levels further support the designations of the new Tunturibacter TAB LE 2 Utilization of various carbon substrates by Tunturibacter species. Utilization of T. empetritectus M8UP23 T. gelidoferens M8UP39 T. psychrotolerans X5P6 T. lichenicola (Belova, Suzina et al., 2018) Acetate    Arabinose     Benzoate    Cellobiose + + + + Dulcitol    Xylose + + + + Fructose + + + + Fucose + + + Galacturonic acid    Glucosamine + + w Glucose + + + + Lactulose    + L-glutamic acid w w w L-glutamine    Maltose + + + + Mannose + + + + Mannitol    w N-acetyl-glucosamine + + w Oxalate    + Pyruvate    Salicin + + + + Sorbitol    Trehalose + + + + Note: +, positive reaction; , negative reaction; w, weak positive reaction. 10 of 19 MESSYASZ ET AL.ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License species and strains (Figure 4). Only the longest assem- bled scaffold for each genome was used for the align- ment. At the genus level we compared T. lichenicola DSM 104462 (5,662,239 bp), G. arctica DSM 23128 (4,736,692 bp), E. modestus DSM 18191 (6,121,180 bp), and Terriglobus saanensis SP1PR4 F I GURE 4 Genome comparisons at various taxonomic levels. (A) Comparison of four SD1 Acidobacteriota including Tunturibacter lichenicola. (B) Comparison of the four species within the Tunturibacter (each species represents the type strain). (C) Comparison of four strains within the Tunturibacter empetritectus. Genome comparisons were viewed via PGV-Mummer. Black horizontal lines indicate the span of the genome being compared, green lines connect homologous regions between the genomes in the same orientation (normal link), and brown lines connect homologous regions between the genomes in reverse orientation (inverted link). The shade of green and brown indicate percent homology (bottom right legend, dark colour means higher percent homology). TUNTURIBACTER GEN. NOV. IN THE ACIDOBACTERIOTA PHYLUM 11 of 19ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License (5,095,225 bp). At the species level we compared T. lichenicola DSM 104462 (5,662,239 bp), T. empetritec- tus M8UP23 (4,898,072 bp), T. gelidoferens M8UP39 (5,453,280 bp), and T. psychrotolerans X5P6 (5,619,635 bp). At the strain level, we also compared T. empetritectus strains M8UP23 (4,898,072 bp), M8UP20 (4,830,045 bp), M8UP22 (5,076,618 bp), and M8UP27 (4,842,676 bp). These alignments show decreased homology when comparing the Tunturibacter genomes against other genera in the Terriglobia (Figure 4A), compared with increased homology when comparing the genomes of the Tunturibacter species against each other (Figure 4B,C). The strength of homology increases with species and strain level alignments, with strain level alignments (within the T. empetritectus strains) showing the least amount of genome rearrangements (Figure 4C). These trends are also apparent when using another genome alignment visualization tool, Mauve (Figure S4A–C), which results in longer homol- ogy segments within alignments at the proposed strains compared with the species and genus level alignments. Unique pathways and phenotypic data of Tunturibacter species The Tunturibacter strains had between 3600 and 7400 predicted genes, of which >98% were protein-coding genes (Table 1). All strains had at least one successful alignment to a gene family in the CAZy HMM database within the following CAZy classes: glycoside hydro- lases, glycosyl transferases, carbohydrate-binding modules, carbohydrate esterases, polysaccharide lyases, and auxiliary activities. For each assembled genome the number of CAZy gene alignments includes one or more hits to unique CAZy genes spanning numerous functions (Table 1 and Figure S5). In the dif- ferent strains, 8%–16% of the total number of predicted genes was coded for modules of the CAZy family, with genes for glycoside hydrolases being the most abun- dant. While these hits represent mostly partial align- ments to CAZy genes, this indicates a wide set of unique genes involved in the build-up and breakdown of complex carbohydrates. Compared with the other Acidobacteriota genomes in the pangenome analysis, several pathways were sig- nificantly enriched (p < 0.05) in members of the Tunturi- bacter genus (Table 3). Based on KEGG Class and COG20 Functional annotations, Tunturibacter genomes are more enriched in functional pathways involving the metabolism of amino acids, such as tryptophan, betaine, and lysine. The enriched GABA (γ-aminobutyric acid) biosynthesis pathway suggests a pH resistance strategy within the Tunturibacter spp. (Dhakal et al., 2012). This genus is also enriched in genes within pathways involv- ing the metabolism and biosynthesis of cofactors and vitamins such as menaquinone, cobalamin, biotin, and molybdopterin. These (COG20 Pathway) annotations do not, however, support the presence of the full pathway within the Tunturibacter. The specific KEGG and COG20 genes annotated from the enriched functional pathways can be found in Table S5. Additionally, there is a lack of evidence for promi- nent vitamin and cofactor biosynthesis genes within many members of the Terriglobia. Menaquinone has been found in Acidobacterium capsulatum (Kishimoto et al., 1991), cobalamin (B12) transport has only been speculated in SD4 and SD8 Acidobacteriota and Bryo- cella species (Kielak et al., 2016). An uncultivated TAB LE 3 Enriched functional pathways within the Tunturibacter species compared with other Acidobacteriota genomes. Annotation Enrichment score Adjusted q-value Accession KEGG module Aromatic amino acid metabolism: Tryptophan metabolism, tryptophan = >kynurenine = >2-aminomuconate; NAD biosynthesis, tryptophan = >quinolinate = >NAD 15.51 0.01 M00038; M00912 Amino acid metabolism: Betaine biosynthesis, choline = > betaine; Lysine degradation, lysine = > saccharopine = > acetoacetyl- CoA; GABA biosynthesis, eukaryotes, putrescine = > GABA 10.79 0.03 M00555; M00032; M00135 COG20 pathway COG20 Pathway: Menaquinone biosynthesis; Lysine biosynthesis 31.02 <0.01 COG0318; COG3320 COG20 Pathway: Cobalamin/B12 biosynthesis 15.51 <0.01 COG1010; COG2243 COG20 Pathway: Pyrimidine salvage; Biotin biosynthesis 10.79 0.02 COG0402; COG2226 COG20 Pathway: Molybdopterin biosynthesis 8.99 0.05 COG0314; COG1977 12 of 19 MESSYASZ ET AL.ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License group of Acidobacteriota, GAL08, from hot springs con- tained a biotin-specific transporter; however, it was not able to synthesize biotin (Ruhl et al., 2022). Although more detailed analysis is needed, the enrichment of vitamin/cofactor biosynthesis genes within the Tunturi- bacter suggests that members of this genus may be able to synthesize vitamins de novo and have increased metabolic capabilities compared with the Edaphobacter, Granulicella, and Terriglobus species. Acidobacteriota community in Malla tundra heath soils We examined the distribution of the Terriglobia in a set of snow-accumulating and windswept tundra heath soils in Malla Nature Reserve, from which the novel Tunturi- bacter species were isolated. At this site, areas in depressions are sheltered from the winds with high snow accumulation (up to ≥1 m), while windswept areas remain essentially snow-free throughout the winter lead- ing to distinctly different soil temperature profiles and dif- ferences in the amplitude of annual temperature variation (Männistö et al., 2024). The bacterial communi- ties were assessed by rRNA operon profiling with the Oxford Nanopore MinION, enabling strain-specific identi- fication of bacterial community members. Overall, rRNA operon reads from the Acidobacteriota represented 4.2%–18.4% of the >760 K total bacterial reads from these tundra samples. Specifically, the Acidobacteriota reads varied from 1524 to 18,818 reads per sample with an average of 6536 ± 5478 per site. Re-screening of these reads against the augmented Acidobacteriota rRNA database, containing the new rRNA operons from this study, indicated that 18%–38% of these Acidobac- teriota reads have best BLAST matches to Tunturibacter and Granulicella. Furthermore, distinct differences in the relative abundance of Edaphobacter, Granulicella, and Terri- globus species/strains could be observed between windswept and snow-accumulating tundra heaths (Figure 5). Reads matching the various Tunturibacter strains were generally more dominant in the windswept plots, while the Granulicella strain was more abundant in the snow-accumulating plots. These strain and spe- cies distinctions correspond to the observed differ- ences in the overall soil microbial communities between windswept and snow-accumulating tundra heaths (Männistö et al., 2024). Snow cover and the linked vegetation shifts and soil C and N dynamics may thus be an important microclimatic driver of bacterial communities. The improved taxonomy of the Acidobac- teriota combined with the genomic and phenotypic information of cultivated strains will enable an improved understanding of the community response to fluctuating environmental conditions in a changing climate. F I GURE 5 Terriglobia community in soils of windswept (K1, K3, K6, K8) and snow-accumulating (T2, T5, T6, T7) tundra heath plots of Mt. Pikku Malla. rRNA operon reads from the Acidobacteriota represent 4.2%–18.4% of the total bacterial reads. TUNTURIBACTER GEN. NOV. IN THE ACIDOBACTERIOTA PHYLUM 13 of 19ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Ecological context of Acidobacteriota in tundra soils Members of the Acidobacteriota are ubiquitous in acidic Arctic tundra and sub-Arctic forest soils. The Kilpisjärvi study site has representative tundra vegetation domi- nated by dwarf shrub-rich Empetrum heaths over acidic soils, or forb- and graminoid-rich Dryas heaths over non-acidic soils (Eskelinen et al., 2009; Männistö et al., 2007, 2009, 2013). These soils are organic-rich and well-aerated, harbouring abundant aerobic hetero- trophic microbiota where diverse Acidobacteriota make up 10%–40% of the total bacterial community (Männistö et al., 2009, 2013). The abundance and com- position of the Acidobacteriota vary with topography and exposure influence snow accumulation and soil temperatures (Männistö et al., 2013) as well as bedrock material which influences soil pH (Männistö et al., 2007). Different species within the Acidobacter- iota appear to respond to environmental conditions dif- ferently, highlighting the wide functional diversity of these organisms even within the Terriglobia. From the tundra heath sites in Kilpisjärvi Finland, several species of Terriglobus, Granulicella, and Tunturibacter have been cultivated (Männistö et al., 2011, 2012, this study). These species appear adapted to the break- down, utilization, and biosynthesis of diverse polysac- charides, and resilience to fluctuating temperatures and nutrient-deficient conditions. Increased competition between plants and microbes and within the microbial community can occur due to environmental factors such as the increased presence of woody evergreen shrubs that immobilize nutrients as a consequence of warming-induced ‘shrubification’ on Arc- tic soil carbon storage (Stark et al., 2023). Tundra soil car- bon and nitrogen cycles are strongly coupled with soil nitrogen availability often decreasing over the growing season, leading to the soil microbial communities becom- ing increasingly N-limited (Jonasson et al., 1999; Stark & Kytöviita, 2006). Methods to combat low nutrient concen- trations include having the functional capacity for nutrient assimilation and transport. Acidobacteriota is known to contain genes involved in attaining nitrogen, which include genes essential for ammonia assimilation, for example, glutamine synthetase, and genes for the ammo- nium channel transporter family (amtB gene; Eichorst et al., 2018). In a recent study comparing microbial responses to Arctic greening in Alaskan soils, Acidobac- teriota most strongly expressed the glutamine synthetase metaprotein, a large component of the proteins involved in ammonia metabolism (Miller et al., 2023). Within our study, glutamine synthetase (COG0174) was annotated mainly within the core genome of the Tunturibacter pan- genome (Figure 3; Bin 1). Another adaptation to low nutrient concentrations includes genes for high-affinity transport systems (Eichorst et al., 2018; Kielak et al., 2016). Transporter genes annotated within the Tunturibacter pangenome include high-affinity Mn2+ porin (COG3659), high- affinity iron transporter (COG4393), and high-affinity nickel/cobalt permease (HoxN; COG3376). Acidobac- teriota has been shown to contain a broad substrate range of transporters, including the Drug/Metabolite transporter superfamily, the Ammonia Transporter Channel (Amt) Family, and the Metal Ion (Mn2+ Iron) Transporter (Nramp; Kielak et al., 2016). This diversity of transporters suggests that Acidobacteriota have an advantage in nutrient assimilation in nutrient-poor conditions. Tundra soil Acidobacteriota appears to be oligo- trophs with slow growth rates and low population turn- over rates that are favoured by the presence of recalcitrant carbon. The genomes of the tundra soil Acidobacteriota contain an abundance of conserved genes/gene clusters encoding for modules of the CAZy families, predominately glycoside hydrolases and gly- cosyltransferases (Figure S5; Eichorst et al., 2018; Rawat et al., 2012). In our study, pangenome glycosyl- transferase COGs involved in cell wall biosynthesis (COG0438) and the glycosylation of proteins and lipid IVA (COG1807) were found across all Acidobacteriota analysed. Multiple alignments to various CAZy gene families were also found within the Tunturibacter indi- cating a wide ability for utilizing complex carbohydrates. Another interesting feature of the Acidobacteriota is their synthesis of hopanoid lipids, which may play a role in regulating membrane stability. Additionally, quorum-sensing associated genes, such as acyl- homoserine lactone (acyl-HSL), were found within several Tunturibacter and across Acidobacteriota within our pangenome analysis (Parsek et al., 1999). Signalling molecules like acyl-HSLs are known to impact bacterial community dynamics in soils, particu- larly in response to virulence genes (Parsek et al., 1999; Riaz et al., 2008). Members of the Acidobacteriota are known to play a role in the decomposition of SOM, in particular various biopolymers, and they participate in the global cycling of carbon, iron, and hydrogen. While Acidobacteriota constitutes from 5% to 50% of the total bacterial com- munity in organic-rich, low pH, Arctic, and boreal soils based on rRNA gene reads (Männistö et al., 2007, 2009, 2013) they continue to be notoriously difficult to cultivate and are thus woefully under-represented in culture collections. The description of new strains and species provides key phenotypic and genomic informa- tion to help us understand why they are so dominant and how they maintain diversity within the group. There is much to learn about their ecosystem functions in soils, their interactions with other microbes, and their adaptations to environmental stress and climate change. 14 of 19 MESSYASZ ET AL.ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Description of Tunturibacter gen. nov. The results from the multiple lines of analyses pre- sented above indicate that E. lichenicola is not a mem- ber of the genus Edaphobacter and should be reclassified as a species of the proposed new genus Tunturibacter, with the proposed name T. lichenicola comb. nov. The NCBI BioProject ID for the 10 newly assembled Tunturibacter genomes is PRJNA1004338. Accession numbers for all new Tunturibacter 16S rRNA genes and genomes are in Table S6. Tunturibacter (Tun’tu.ri.bac’ter. tunturi, referring to the Arctic treeless fells in Finland, N.L. masc. n. bacter, a short rod; N.L. masc. n. Tunturibacter rod-shaped tundra fell bacterium). Cells are Gram-negative aerobic rods. On agar plates, cells form circular, mucoid colonies. Colony pig- ment varies from light beige to pink. Produce extracellu- lar polysaccharide-like substances. Sugars are preferred carbon sources and the different strains are capable of hydrolyzing various polysaccharides. The major cellular fatty acids are iso-15:0, iso-13:0, 16:1ω7c, and 16:0. The predominant menaquinone is MK-8. The DNA G + C content is 56%–58%. Strains have been isolated from tundra soil and lichen thalli col- lected from lichen-dominated forested tundra. The type species is T. lichenicola comb. nov. Emended description of T. lichenicola comb. nov. The description is as given by Belova, Ravin, et al. (2018) and Belova, Suzina, et al. (2018). The type strain is SBC68T (=DSM 104462T = VKM B-3208T). Description of Tunturibacter empetritectus sp. nov. Tunturibacter empetritectus (em.pet.ri.tectus. empetri, referring to the plant Empetrum nigrum ssp. hermaph- roditum; L. adj. tectus, covered; L. part. Adj. empetritec- tus, growing under tundra heath dominated by E. nigrum ssp. hermaphroditum). Cells are Gram-negative, non-motile, aerobic rods. Colonies are light beige or pink, circular, and smooth when grown on GY agar (Figure S6). Growth occurs at 2–30C and pH 3.5–6.5. In VL55 mineral medium with 100 mg L1 yeast extract and 5–10 mM carbon source, utilizes cellobiose, xylose, fructose, fucose, glucos- amine, glucose, L-glutamic acid (weak), maltose, man- nose, N-acetyl-glucosamine, salicin and trehalose. Hydrolyzes xylan, pectin, xanthan and gum arabic. The major cellular fatty acids are iso-15:0, iso-13:0, 16:1ω7c, and 16:0. The DNA G + C content determined from the genome sequence of the type strain is 57.82%. The type strain is M8UP23T (= DSMZ 117310 = HAMBI 3810) isolated from tundra soil in Malla Nature Reserve, Kilpisjärvi, Finland (69010 N, 20500 E). NCBI accession numbers for the 16S rRNA gene sequence and the draft genome sequence of the type strain are OR449309 and CP132932-CP132934, respectively. Description of T. gelidoferens sp. nov. Tunturibacter gelidoferens (ge.li.do.ferens. L. adj. gelido cold; L. pres. Part. Ferens, to endure; L. part. Adj. gelidoferens, cold-enduring). Cells are Gram-negative, non-motile, aerobic rods. Colonies are pink, circular, and smooth when grown on GY agar (Figure S6). Growth occurs at 2–30C and pH 3.5–6.5. In VL55 mineral medium with 100 mg L1 yeast extract and 5–10 mM carbon source, utilizes cellobiose, xylose, fructose, fucose, glucosamine, glu- cose, L-glutamic acid (weak), maltose, mannose, N- acetyl-glucosamine, salicin, and trehalose. Hydrolyzes carboxy methyl cellulose (weak), xylan, pectin, lichenin, starch (weak), and gum arabic (weak). The major cellu- lar fatty acids are iso-15:0, iso-13:0, 16:1ω7c, and 16:0. The DNA G + C content determined from the genome sequence of the type strain is 57.42%. The type strain is M8UP39T (= DSMZ 117311 = HAMBI 3809) isolated from tundra soil in Malla Nature Reserve, Kilpisjärvi, Finland (69010 N, 20500 E). NCBI accession numbers for the 16S rRNA gene sequence and the draft genome sequence of the type strain are OR449311 and CP132937–CP132938, respectively. Description of T. psychrotolerans sp. nov. Tunturibacter psychrotolerans (psy.chro.to’le.rans. Gr. Adj. psychros, cold; L. pres. Part. Tolerans, tolerat- ing; N.L. part. Adj. psychrotolerans, cold-tolerating). Cells are Gram-negative, non-motile, aerobic rods. Colonies are light beige to light pink, circular, and smooth when grown on GY agar (Figure S6). Growth occurs at 2–30C and pH 3.5–6.5. In VL55 mineral medium with 100 mg L1 yeast extract and 5–10 mM carbon source, utilizes cellobiose, xylose, fructose, fucose, glucosamine (weak), glucose, L-glutamic acid (weak), maltose, mannose, N-acetyl-glucosamine (weak), salicin, and trehalose. Hydrolyzes carboxy methyl cellulose (weak), pectin, lichenin, starch, xanthan and gum arabic (weak). The major cellular fatty acids are iso-15:0, iso-13:0, 16:1ω7c, and 16:0. The DNA G + C content determined from the genome sequence of the type strain is 56.72%. TUNTURIBACTER GEN. NOV. IN THE ACIDOBACTERIOTA PHYLUM 15 of 19ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License The type strain is X5P6T (= DSMZ 117309 = HAMBI 3811) isolated from tundra soil in Malla Nature Reserve, Kilpisjärvi, Finland (69010 N, 20500 E). NCBI accession numbers for the 16S rRNA gene sequence and the draft genome sequence of the type strain are OR449310 and CP132942–CP132943, respectively. AUTHOR CONTRIBUTIONS Adriana Messyasz: Writing—original draft; formal analysis; investigation; writing—review and editing; visualization; data curation; methodology. Minna K. Männistö: Investigation; writing—review and edit- ing; funding acquisition; conceptualization; methodol- ogy. Lee J. Kerkhof: Writing—review and editing; supervision; funding acquisition; resources; methodol- ogy; investigation. Max M. Häggblom: Supervision; writing—review and editing; funding acquisition; con- ceptualization; project administration; methodology; investigation; resources. ACKNOWLEDGEMENTS We thank Sirkka-Liisa Aakkonen, Sari Välitalo, and Marika Pätsi for their help in cultivating and characteriz- ing the strains. We thank Kristina Chew and Serena Connolly for their assistance with nomenclature. This study was funded in part by the US National Science Foundation (Award Number 2129351) to MMH and LJK, the Academy of Finland (decision numbers 130507 and 310776) to MKM, and the USDA National Institute of Food and Agriculture Hatch project acces- sion number 1012785 through the New Jersey Agricul- tural Experiment Station (Hatch Project NJ01160) to MMH. Illumina sequencing of select strains was done through the Community Science Program (CSP) of the US Department of Energy Joint Genomes Institute (Genomic Sequencing of Core and Pangenomes of Soil and Plant-associated Prokaryotes; PI William B. Whitman). CONFLICT OF INTEREST STATEMENT The authors declare no conflicts of interest. DATA AVAILABILITY STATEMENT The NCBI BioProject ID for the newly assembled Tun- turibacter genomes is PRJNA1004338: https://www. ncbi.nlm.nih.gov/bioproject/PRJNA1004338. Accession numbers for 16S rRNA genes are OR449309– OR449318. Accession numbers for genomes are CP132926-CP132945. The rRNA operon reads from Malla Nature Reserve soil samples are available in Bio- Project ID PRJNA1093128: https://www.ncbi.nlm.nih. gov/bioproject/PRJNA1093128. ORCID Max M. Häggblom https://orcid.org/0000-0001-6307- 7863 REFERENCES Barns, S.M., Cain, E.C., Sommerville, L. & Kuske, C.R. (2007) Acido- bacteria phylum sequences in uranium-contaminated subsur- face sediments greatly expand the known diversity within the phylum. Applied and Environmental Microbiology, 73, 3113– 3116. Belova, S.E., Ravin, N.V., Pankratov, T.A., Rakitin, A.L., Ivanova, A.A., Beletsky, A.V. et al. (2018) Hydrolytic capabilities as a key to environmental success: chitinolytic and cellulolytic Acidobacteria from acidic sub-arctic soils and boreal peatlands. Frontiers in Microbiology, 9, 2775. Belova, S.E., Suzina, N.E., Rijpstra, W.I.C., Sinninghe Damsté, J.S. & Dedysh, S.N. (2018) Edaphobacter lichenicola sp. nov., a mem- ber of the family Acidobacteriaceae from lichen-dominated for- ested tundra. International Journal of Systematic and Evolutionary Microbiology, 68, 1265–1270. Bond-Lamberty, B. & Thomson, A. (2010) Temperature-associated increases in the global soil respiration record. Nature, 464, 579–582. Campbell, B.J., Polson, S.W., Hanson, T.E., Mack, M.C. & Schuur, E.A.G. (2010) The effect of nutrient deposition on bacte- rial communities in Arctic tundra soil. Environmental Microbiol- ogy, 12, 1842–1854. Challacombe, J.F., Eichorst, S.A., Hauser, L., Land, M., Xie, G. & Kuske, C.R. (2011) Biological consequences of ancient gene acquisition and duplication in the large genome of Candidatus Solibacter usitatus Ellin6076. PLoS One, 6, e24882. Chesson, P. (2000) Mechanisms of maintenance of species diversity. Annual Review of Ecology and Systematics, 31, 343–366. Chesson, P. & Huntly, N. (1997) The roles of harsh and fluctuating conditions in the dynamics of ecological communities. The American Naturalist, 150, 519–553. Chu, H., Fierer, N., Lauber, C.L., Caporaso, J.G., Knight, R. & Grogan, P. (2010) Soil bacterial diversity in the Arctic is not fun- damentally different from that found in other biomes. Environ- mental Microbiology, 12, 2998–3006. Crits-Christoph, A., Diamond, S., Al-Shayeb, B., Valentin- Alvarado, L. & Banfield, J.F. (2022) A widely distributed genus of soil Acidobacteria genomically enriched in biosynthetic gene clusters. ISME Communications, 2, 1–8. Darling, A.C.E., Mau, B., Blattner, F.R. & Perna, N.T. (2004) Mauve: multiple alignment of conserved genomic sequence with rearran- gements. Genome Research, 14, 1394–1403. Davis, K.E.R., Joseph, S.J. & Janssen, P.H. (2005) Effects of growth medium, inoculum size, and incubation time on culturability and isolation of soil bacteria. Applied and Environmental Microbiol- ogy, 71, 826–834. Dedysh, S.N., Ivanova, A.A., Begmatov, S.A., Beletsky, A.V., Rakitin, A.L., Mardanov, A.V. et al. (2022) Acidobacteria in fens: phylogenetic diversity and genome analysis of the key represen- tatives. Microbiology, 91, 662–670. Dedysh, S.N. & Yilmaz, P. (2018) Refining the taxonomic structure of the phylum Acidobacteria. International Journal of Systematic and Evolutionary Microbiology, 68, 3796–3806. Dhakal, R., Bajpai, V.K. & Baek, K.-H. (2012) Production of gaba (γ – aminobutyric acid) by microorganisms: a review. Brazilian Jour- nal of Microbiology, 43, 1230–1241. Diamond, S., Andeer, P.F., Li, Z., Crits-Christoph, A., Burstein, D., Anantharaman, K. et al. (2019) Mediterranean grassland soil C– N compound turnover is dependent on rainfall and depth, and is mediated by genomically divergent microorganisms. Nature Microbiology, 4, 1356–1367. Eddy, S.R. (2011) Accelerated profile HMM searches. PLoS Compu- tational Biology, 7, e1002195. Edgar, R.C. (2021) MUSCLE v5 enables improved estimates of phy- logenetic tree confidence by ensemble bootstrapping. 2021.06.20.449169. 16 of 19 MESSYASZ ET AL.ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Eichorst, S.A., Breznak, J.A. & Schmidt, T.M. (2007) Isolation and characterization of soil bacteria that define Terriglobus gen. nov., in the phylum Acidobacteria. Applied and Environmental Microbiology, 73, 2708–2717. Eichorst, S.A., Trojan, D., Roux, S., Herbold, C., Rattei, T. & Woebken, D. (2018) Genomic insights into the Acidobacteria reveal strategies for their success in terrestrial environments. Environmental Microbiology, 20, 1041–1063. Emms, D.M. & Kelly, S. (2019) OrthoFinder: phylogenetic orthology inference for comparative genomics. Genome Biology, 20, 238. Eren, A.M., Kiefl, E., Shaiber, A., Veseli, I., Miller, S.E., Schechter, M.S. et al. (2021) Community-led, integrated, repro- ducible multi-omics with anvi’o. Nature Microbiology, 6, 3–6. Eskelinen, A., Stark, S. & Männistö, M. (2009) Links between plant com- munity composition, soil organic matter quality and microbial com- munities in contrasting tundra habitats. Oecologia, 161, 113–123. Fry, E.L., Ashworth, D., Allen, K.A.J., Chardon, N.I., Rixen, C., Björkman, M.P. et al. (2023) Vegetation type, not the legacy of warming, modifies the response of microbial functional genes and greenhouse gas fluxes to drought in oro-arctic and alpine regions. FEMS Microbiology Ecology, 99, fiad145. Gadkari, P.S., McGuinness, L.R., Männistö, M.K., Kerkhof, L.J. & Häggblom, M.M. (2020) Arctic tundra soil bacterial communities active at subzero temperatures detected by stable isotope prob- ing. FEMS Microbiology Ecology, 96, fiz192. German Collection of Microorganisms and Cell Cultures GmbH: Polar Lipids. (n.d.) German Collection of Microorganisms and Cell Cultures GmbH: Respiratory Quinones. (n.d.) Ghoul, M. & Mitri, S. (2016) The ecology and evolution of microbial competition. Trends in Microbiology, 24, 833–845. Göker, M. (2023) Filling the gaps: missing taxon names at the ranks of class, order and family. International Journal of Systematic and Evolutionary Microbiology, 72, 005638. Graham, D.E., Wallenstein, M.D., Vishnivetskaya, T.A., Waldrop, M.P., Phelps, T.J., Pfiffner, S.M. et al. (2012) Microbes in thawing permafrost: the unknown variable in the climate change equation. The ISME Journal, 6, 709–712. Gurevich, A., Saveliev, V., Vyahhi, N. & Tesler, G. (2013) QUAST: quality assessment tool for genome assemblies. Bioinformatics, 29, 1072–1075. Hugelius, G., Strauss, J., Zubrzycki, S., Harden, J.W., Schuur, E.A.G., Ping, C.-L. et al. (2014) Estimated stocks of cir- cumpolar permafrost carbon with quantified uncertainty ranges and identified data gaps. Biogeosciences, 11, 6573–6593. Ivanova, A.A., Zhelezova, A.D., Chernov, T.I. & Dedysh, S.N. (2020) Linking ecology and systematics of acidobacteria: distinct habi- tat preferences of the Acidobacteriia and Blastocatellia in tundra soils. PLoS One, 15, e0230157. Janssen, P.H. (2006) Identifying the dominant soil bacterial taxa in libraries of 16S rRNA and 16S rRNA genes. Applied and Envi- ronmental Microbiology, 72, 1719–1728. Jonasson, S., Michelsen, A. & Schmidt, I.K. (1999) Coupling of nutri- ent cycling and carbon dynamics in the Arctic, integration of soil microbial and plant processes. Applied Soil Ecology, 11, 135–146. Jones, R.T., Robeson, M.S., Lauber, C.L., Hamady, M., Knight, R. & Fierer, N. (2009) A comprehensive survey of soil acidobacterial diversity using pyrosequencing and clone library analyses. The ISME Journal, 3, 442–453. Kalam, S., Basu, A., Ahmad, I., Sayyed, R.Z., El-Enshasy, H.A., Dailin, D.J. et al. (2020) Recent understanding of soil Acidobac- teria and their ecological significance: a critical review. Frontiers in Microbiology, 11, 580024. Kerkhof, L.J., Dillon, K.P., Häggblom, M.M. & McGuinness, L.R. (2017) Profiling bacterial communities by MinION sequencing of ribosomal operons.Microbiome, 5, 116. Kerkhof, L.J., Roth, P.A., Deshpande, S.V., Bernhards, R.C., Liem, A. T., Hill, J.M. et al. (2022) A ribosomal operon database and MegaBLAST settings for strain-level resolution of microbiomes. FEMS Microbes, 3, xtac002. Kielak, A.M., Barreto, C.C., Kowalchuk, G.A., van Veen, J.A. & Kuramae, E.E. (2016) The ecology of Acidobacteria: moving beyond genes and genomes. Frontiers in Microbiology, 7, 744. Kim, H.M., Jung, J.Y., Yergeau, E., Hwang, C.Y., Hinzman, L., Nam, S. et al. (2014) Bacterial community structure and soil properties of a subarctic tundra soil in council, Alaska. FEMS Microbiology Ecology, 89, 465–475. Kishimoto, N., Kosako, Y. & Tano, T. (1991) Acidobacterium capsula- tum gen. Nov., sp. nov.: an acidophilic chemoorganotrophic bac- terium containing menaquinone from acidic mineral environment. Current Microbiology, 22, 1–7. Koch, I.H., Gich, F., Dunfield, P.F. & Overmann, J. (2008) Edapho- bacter modestus gen. nov., sp. nov., and Edaphobacter aggre- gans sp. nov., acidobacteria isolated from alpine and forest soils. International Journal of Systematic and Evolutionary Microbiology, 58, 1114–1122. Kolmogorov, M., Yuan, J., Lin, Y. & Pevzner, P.A. (2019) Assembly of long, error-prone reads using repeat graphs. Nature Biotechnol- ogy, 37, 540–546. Koven, C.D., Ringeval, B., Friedlingstein, P., Ciais, P., Cadule, P., Khvorostyanov, D. et al. (2011) Permafrost carbon-climate feed- backs accelerate global warming. Proceedings of the National Academy of Sciences of the United States of America, 108, 14769–14774. Lane, D.J. (1991) 16S/23S rRNA sequencing. In: Stackebrandt, E. & Goodfellow, M. (Eds.) Nucleic acid techniques in bacterial sys- tematics. Chichester: Wiley, pp. 115–175. Lee, M.D. (2019) GtoTree: a user-friendly workflow for phyloge- nomics. Bioinformatics, 35, 4162–4164. Lennon, J.T., Aanderud, Z.T., Lehmkuhl, B.K. & Schoolmaster, D.R., Jr. (2012) Mapping the niche space of soil microorganisms using taxonomy and traits. Ecology, 93, 1867–1879. Letunic, I. & Bork, P. (2021) Interactive tree of life (iTOL) v5: an online tool for phylogenetic tree display and annotation. Nucleic Acids Research, 49, W293–W296. Lin, Y., Yuan, J., Kolmogorov, M., Shen, M.W., Chaisson, M. & Pevzner, P.A. (2016) Assembly of long error-prone reads using de Bruijn graphs. Proceedings of the National Academy of Sci- ences of the United States of America, 113, E8396–E8405. Loya, W.M. & Grogan, P. (2004) Carbon conundrum on the tundra. Nature, 431, 406–408. Männistö, M., Vuosku, J., Stark, S., Saravesi, K., Suokas, M., Markkola, A. et al. (2018) Bacterial and fungal communities in boreal forest soil are insensitive to changes in snow cover condi- tions. FEMS Microbiology Ecology, 94, fiy123. Männistö, M.K., Ahonen, S.H.K., Ganzert, L., Tiirola, M., Stark, S. & Häggblom, M.M. (2024) Bacterial and fungal communities in sub-Arctic tundra heaths are shaped by contrasting snow accu- mulation and nutrient availability. FEMS Microbiology Ecology, 100, fiae036. Männistö, M.K., Kurhela, E., Tiirola, M. & Häggblom, M.M. (2013) Acidobacteria dominate the active bacterial communities of Arc- tic tundra with widely divergent winter-time snow accumulation and soil temperatures. FEMS Microbiology Ecology, 84, 47–59. Männistö, M.K., Rawat, S., Starovoytov, V. & Häggblom, M.M. (2011) Terriglobus saanensis sp. nov., an acidobacterium isolated from tundra soil. International Journal of Systematic and Evolutionary Microbiology, 61, 1823–1828. Männistö, M.K., Rawat, S., Starovoytov, V. & Häggblom, M.M. (2012) Granulicella arctica sp. nov., Granulicella mallensis sp. nov., Gran- ulicella tundricola sp. nov. and Granulicella sapmiensis sp. nov., novel acidobacteria from tundra soil. International Journal of Sys- tematic and Evolutionary Microbiology, 62, 2097–2106. TUNTURIBACTER GEN. NOV. IN THE ACIDOBACTERIOTA PHYLUM 17 of 19ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Männistö, M.K., Tiirola, M. & Häggblom, M.M. (2007) Bacterial com- munities in Arctic fjelds of Finnish Lapland are stable but highly pH-dependent. FEMS Microbiology Ecology, 59, 452–465. Männistö, M.K., Tiirola, M. & Häggblom, M.M. (2009) Effect of freeze- thaw cycles on bacterial communities of arctic tundra soil. Micro- bial Ecology, 58, 621–631. McMahon, S.K., Wallenstein, M.D. & Schimel, J.P. (2011) A cross- seasonal comparison of active and total bacterial community composition in Arctic tundra soil using bromodeoxyuridine label- ing. Soil Biology and Biochemistry, 43, 287–295. Miller, S.E., Colman, A.S. & Waldbauer, J.R. (2023) Metaproteomics reveals functional partitioning and vegetational variation among permafrost-affected Arctic soil bacterial communities. mSys- tems, 8, e01238-22. Natali, S.M., Schuur, E.A.G., Webb, E.E., Pries, C.E.H. & Crummer, K.G. (2014) Permafrost degradation stimulates car- bon loss from experimentally warmed tundra. Ecology, 95, 602–608. Nikrad, M.P., Kerkhof, L.J. & Häggblom, M.M. (2016) The subzero microbiome: microbial activity in frozen and thawing soils. FEMS Microbiology Ecology, 92, fiw081. Oechel, W.C., Vourlitis, G. & Hastings, S.J. (1997) Cold season CO2 emission from Arctic soils. Global Biogeochemical Cycles, 11, 163–172. Ondov, B.D., Treangen, T.J., Melsted, P., Mallonee, A.B., Bergman, N.H., Koren, S. et al. (2016) Mash: fast genome and metagenome distance estimation using MinHash. Genome Biol- ogy, 17, 132. Pankratov, T.A. & Dedysh, S.N. (2010) Granulicella paludicola gen. nov., sp. nov., Granulicella pectinivorans sp. nov., Granulicella aggregans sp. nov. and Granulicella rosea sp. nov., acidophilic, polymer-degrading acidobacteria from Sphagnum peat bogs. International Journal of Systematic and Evolutionary Microbiol- ogy, 60, 2951–2959. Parks, D.H., Chuvochina, M., Rinke, C., Mussig, A.J., Chaumeil, P.-A. & Hugenholtz, P. (2022) GTDB: an ongoing census of bacterial and archaeal diversity through a phylogeneti- cally consistent, rank normalized and complete genome-based taxonomy. Nucleic Acids Research, 50, D785–D794. Parks, D.H., Imelfort, M., Skennerton, C.T., Hugenholtz, P. & Tyson, G.W. (2015) CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagen- omes. Genome Research, 25, 1043–1055. Parsek, M.R., Val, D.L., Hanzelka, B.L., Cronan, J.E. & Greenberg, E.P. (1999) Acyl homoserine-lactone quorum-sens- ing signal generation. Proceedings of the National Academy of Sciences of the United States of America, 96, 4360–4365. Parsley, L.C., Linneman, J., Goode, A.M., Becklund, K., George, I., Goodman, R.M. et al. (2011) Polyketide synthase pathways identified from a metagenomic library are derived from soil Acid- obacteria. FEMS Microbiology Ecology, 78, 176–187. Pessi, I.S., Viitamäki, S., Virkkala, A.-M., Eronen-Rasimus, E., Delmont, T.O., Marushchak, M.E. et al. (2022) In-depth charac- terization of denitrifier communities across different soil ecosys- tems in the tundra. Environmental Microbiomes, 17, 30. Podar, M., Turner, J., Burdick, L.H. & Pelletier, D.A. (2019) Complete genome sequence of Terriglobus albidus strain ORNL, an Acido- bacterium isolated from the Populus deltoides rhizosphere. Microbiology Resource Announcements, 8, e01065-19. Poppeliers, S.W.M., Hefting, M., Dorrepaal, E. & Weedon, J.T. (2022) Functional microbial ecology in arctic soils: the need for a year- round perspective. FEMS Microbiology Ecology, 98, fiac134. Price, M.N., Dehal, P.S. & Arkin, A.P. (2010) FastTree 2 – approxi- mately maximum-likelihood trees for large alignments. PLoS One, 5, e9490. Pritchard, L., Glover, R.H., Humphris, S., Elphinstone, J.G. & Toth, I.K. (2015) Genomics and taxonomy in diagnostics for food security: soft-rotting enterobacterial plant pathogens. Analytical Methods, 8, 12–24. Rawat, S.R., Männistö, M.K., Starovoytov, V., Goodwin, L., Nolan, M., Hauser, L. et al. (2012) Complete genome sequence of Terriglobus saanensis type strain SP1PR4(T), an Acidobac- teria from tundra soil. Standards in Genomic Sciences, 7, 59–69. Rawat, S.R., Männistö, M.K., Starovoytov, V., Goodwin, L., Nolan, M., Hauser, L. et al. (2014) Complete genome sequence of Granulicella tundricola type strain MP5ACTX9(T), an Acido- bacteria from tundra soil. Standards in Genomic Sciences, 9, 449–461. Rawat, S.R., Männistö, M.K., Starovoytov, V., Goodwin, L., Nolan, M., Hauser, L.J. et al. (2013) Complete genome sequence of Granulicella mallensis type strain MP5ACTX8T, an acidobacterium from tundra soil. Standards in Genomic Sci- ences, 9, 71–82. Riaz, K., Elmerich, C., Moreira, D., Raffoux, A., Dessaux, Y. & Faure, D. (2008) A metagenomic analysis of soil bacteria extends the diversity of quorum-quenching lactonases. Environ- mental Microbiology, 10, 560–570. Ruhl, I.A., Sheremet, A., Furgason, C.C., Krause, S., Bowers, R.M., Jarett, J.K. et al. (2022) GAL08, an uncultivated group of Acido- bacteria, is a dominant bacterial clade in a neutral hot spring. Frontiers in Microbiology, 12, 787651. Schmartz, G.P., Hartung, A., Hirsch, P., Kern, F., Fehlmann, T., Müller, R. et al. (2021) PLSDB: advancing a comprehensive database of bacterial plasmids. Nucleic Acids Research, 50, D273–D278. Schuur, E.A.G., Bockheim, J., Canadell, J.G., Euskirchen, E., Field, C.B., Goryachkin, S.V. et al. (2008) Vulnerability of perma- frost carbon to climate change: implications for the global carbon cycle. Bioscience, 58, 701–714. Seeman, T. (2014) Prokka: rapid prokaryotic genome annotation. Bio- informatics, 15, 2068–2069. Shaiber, A., Willis, A.D., Delmont, T.O., Roux, S., Chen, L.-X., Schmid, A.C. et al. (2020) Functional and genetic markers of niche partitioning among enigmatic members of the human oral microbiome. Genome Biology, 21, 292. Shimoyama, Y. (2022) pyGenomeViz: A genome visualization python package for comparative genomics. pyGenomeViz Stark, S., Kumar, M., Myrsky, E., Vuorinen, J., Kantola, A.M., Telkki, V.-V. et al. (2023) Decreased soil microbial nitrogen under vegetation ‘shrubification’ in the subarctic forest–tundra ecotone: the potential role of increasing nutrient competition between plants and soil microorganisms. Ecosystems, 26, 1504–1523. Stark, S. & Kytöviita, M.-M. (2006) Simulated grazer effects on micro- bial respiration in a subarctic meadow: implications for nutrient competition between plants and soil microorganisms. Applied Soil Ecology, 31, 20–31. Tas¸, N., Prestat, E., Wang, S., Wu, Y., Ulrich, C., Kneafsey, T. et al. (2018) Landscape topography structures the soil microbiome in arctic polygonal tundra. Nature Communications, 9, 777. Tveit, A.T., Urich, T. & Svenning, M.M. (2014) Metatranscriptomic analysis of arctic peat soil microbiota. Applied and Environmen- tal Microbiology, 80, 5761–5772. Vaser, R. & Sˇikic, M. (2021) Time- and memory-efficient genome assembly with raven. Nature Computational Science, 1, 332–336. Viitamäki, S., Pessi, I.S., Virkkala, A.-M., Niittynen, P., Kemppinen, J., Eronen-Rasimus, E. et al. (2022) The activity and functions of soil microbial communities in the Finnish sub-Arctic vary across vegetation types. FEMS Microbiology Ecology, 98, fiac079. Vorˇísˇkova, J., Elberling, B. & Priemé, A. (2019) Fast response of fungal and prokaryotic communities to climate change manipulation in two contrasting tundra soils. Environmental Microbiomes, 14, 6. 18 of 19 MESSYASZ ET AL.ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Wang, J., Chen, M.-H., Lv, Y.-Y., Jiang, Y.-W. & Qiu, L.-H. (2016) Edaphobacter dinghuensis sp. nov., an acidobacterium isolated from lower subtropical forest soil. International Journal of Sys- tematic and Evolutionary Microbiology, 66, 276–282. Ward, N.L., Challacombe, J.F., Janssen, P.H., Henrissat, B., Coutinho, P.M., Wu, M. et al. (2009) Three genomes from the phylum Acidobacteria provide insight into the lifestyles of these microorganisms in soils. Applied and Environmental Microbiol- ogy, 75, 2046–2056. Weintraub, M.N. & Schimel, J.P. (2003) Interactions between carbon and nitrogen mineralization and soil organic matter chemistry in arctic tundra soils. Ecosystems, 6, 129–143. Welker, J.M., Fahnestock, J.T. & Jones, M.H. (2000) Annual CO2 flux in dry and moist arctic tundra: Field responses to increases in summer temperatures and winter snow depth. Climatic Change, 44, 139–150. Wick, R.R. & Holt, K.E. (2019) Benchmarking of long-read assem- blers for prokaryote whole genome sequencing. F1000Res, 8, 2138. Wick, R.R. & Holt, K.E. (2022) Polypolish: short-read polishing of long-read bacterial genome assemblies. PLoS Computational Biology, 18, e1009802. Wick, R.R., Judd, L.M., Cerdeira, L.T., Hawkey, J., Méric, G., Vezina, B. et al. (2021) Trycycler: consensus long-read assem- blies for bacterial genomes. Genome Biology, 22, 266. Yoon, S.H., Ha, S.M., Kwon, S., Lim, J., Kim, Y., Seo, H. et al. (2017) Introducing EzBioCloud: a taxonomically united data- base of 16S rRNA and whole genome assemblies. Interna- tional Journal of Systematic and Evolutionary Microbiology, 67, 1613–1617. Zhang, H., Yohe, T., Huang, L., Entwistle, S., Wu, P., Yang, Z. et al. (2018) dbCAN2: a meta server for automated carbohydrate-active enzyme annotation. Nucleic Acids Research, 46(W1), W95–W101. Zhang, Q.-M., Fu, J.-C., Chen, Z.-Q. & Qiu, L.-H. (2022) Paracidobac- terium acidisoli gen. nov., sp. nov. and Alloacidobacterium din- ghuense gen. nov., sp. nov., two acidobacteria isolated from forest soil, and reclassification of Acidobacterium ailaaui and Acidipila dinghuensis as Pseudacidobacterium ailaaui gen. nov., comb. nov. and Silvibacterium dinghuense comb. nov. Int J Syst Evol Microbiol, 72, 005415. Zinger, L., Shahnavaz, B., Baptist, F., Geremia, R.A. & Choler, P. (2009) Microbial diversity in alpine tundra soils correlates with snow cover dynamics. The ISME Journal, 3, 850–859. SUPPORTING INFORMATION Additional supporting information can be found online in the Supporting Information section at the end of this article. How to cite this article: Messyasz, A., Männistö, M.K., Kerkhof, L.J. & Häggblom, M.M. (2024) Genome analysis and description of Tunturibacter gen. nov. expands the diversity of Terriglobia in tundra soils. Environmental Microbiology, 26(5), e16640. Available from: https://doi.org/10.1111/1462-2920.16640 TUNTURIBACTER GEN. NOV. IN THE ACIDOBACTERIOTA PHYLUM 19 of 19ENVIRONMENTAL MICROBIOLOGY 14622920, 2024, 5, D ow nloaded from https://envirom icro-journals.onlinelibrary.wiley.com/doi/10.1111/1462-2920.16640 by Luonnonvarakeskus, W iley Online Library on [27/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on W iley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License