Skip to main content

De novo genome assembly resolving repetitive structures enables genomic analysis of 35 European Mycoplasmopsis bovis strains

Abstract

Mycoplasmopsis (M.) bovis, the agent of mastitis, pneumonia, and arthritis in cattle, harbors a small genome of approximately 1 Mbp. Combining data from Illumina and Nanopore technologies, we sequenced and assembled the genomes of 35 European strains and isolate DL422_88 from Cuba. While the high proportion of repetitive structures in M. bovis genomes represent a particular challenge, implementation of our own pipeline Mycovista (available on GitHub www.github.com/sandraTriebel/mycovista) in a hybrid approach enabled contiguous assembly of the genomes and, consequently, improved annotation rates considerably. To put our European strain panel in a global context, we analyzed the new genome sequences together with 175 genome assemblies from public databases. Construction of a phylogenetic tree based on core genes of these 219 strains revealed a clustering pattern according to geographical origin, with European isolates positioned on clades 4 and 5. Genomic data allowing assignment of strains to tissue specificity or certain disease manifestations could not be identified. Seven strains isolated from cattle with systemic circular condition (SCC), still a largely unknown manifestation of M. bovis disease, were located on both clades 4 and 5. Pairwise association analysis revealed 108 genomic elements associated with a particular clade of the phylogenetic tree. Further analyzing these hits, 25 genes are functionally annotated and could be linked to a M. bovis protein, e.g. various proteases and nucleases, as well as ten variable surface lipoproteins (Vsps) and other surface proteins. These clade-specific genes could serve as useful markers in epidemiological and clinical surveys.

Peer Review reports

Introduction

Infections of the bovine pathogen Mycoplasmopsis (M.) bovis are of considerable economic importance due to their negative impact on animal health and production yields [1]. Pneumonia and mastitis are the most prominent clinical manifestations in cattle, but arthritis, genital disorders, and keratoconjunctivitis can also be caused by this agent [2,3,4]. Occasionally, M. bovis infections presenting as purulent fibrinous pleuropneumonia with sequestering were observed in adult cattle and had to be differentiated from other Mycoplasmopsis diseases, such as Contagious Bovine Pleuropneumonia. Recently, an increase of acute M. bovis infections leading to mastitis in medium-sized and large dairy herds was observed in the eastern and northern federal states of Germany. In some cases, a different clinical picture compared to previous outbreaks was noticed since, in addition to mastitis, affected animals showed symptoms of circulatory involvement manifesting as massive edema in the chest, abdomen, and leg area, similar to allergic reactions (see Fig. S1) [5, 6]. We will refer to this disease complex as Systemic Circular Condition (SCC).

While many aspects of M. bovis pathogenesis are still not fully understood, the agent’s capability of subverting host immune response through surface antigen variation was characterized in several studies [7,8,9,10]. The central role in this process is played by members of a family of lipoproteins designated variable surface proteins (Vsps), which consist of a conserved N-terminal domain for membrane insertion and lipoprotein processing and a large stretch of repetitive tandem domains comprising up to 80 % of the entire Vsp molecule. As a result of spontaneous and non-coordinated deletions, insertions, and rearrangements in the vsp genomic locus, translated Vsps undergo variations in phase (on/off switching), size (varying number of tandem repeats), and/or surface presentation at high frequency [8].

In this context, sequencing and comparative genomics can be important tools to identify genetic features correlating with strain properties and/or disease symptoms [11]. High-quality genomes with precise and, if possible, complete annotations are needed to perform meaningful comparative studies. However, despite their small size (\(\sim\) 1 Mbp), the assembly of M. bovis genomes is challenging due to the high rate of the above-mentioned repetitive structures. In the past, using state-of-the-art tools to assemble Illumina reads resulted in fragmented genomes because repetitive regions are not efficiently covered using short reads. Nevertheless, despite the lower contiguity of such short-read assemblies, the high sequence accuracy of short reads and the associated accurate annotation of open reading frames (ORFs) is essential for further downstream analysis. Currently, 588 assemblies of M. bovis strains are available in the NCBI database (Jan 16, 2023), of which 51 assemblies are marked as complete. Nine of them lack GenBank and/or RefSeq annotation entries. The current NCBI reference genome of M. bovis (strain 8790, GCF_005061465.1) consists of 17 contigs and is incomplete. Sequencing via nanopores, e.g. the Oxford Nanopore Technologies (ONT) MinION device, has great potential to enable genome assemblies of higher contiguity since longer reads facilitate coverage of complex genomic regions [12, 13]. Recent studies have shown that the use of long and short reads is a suitable strategy for assembling bacterial genomes [14].

In the present study, we used a hybrid assembly approach combining the advantages of Illumina and Nanopore sequencing to obtain high-quality genome sequences of 36 M. bovis strains. The strain panel contains seven isolates associated with the SCC disease complex in Germany from the last ten years, strains from different geographical regions and from animals with other symptoms, as well as previous isolates. Genome sequence data were processed using our own Mycovista pipeline (www.github.com/sandraTriebel/mycovista), which also includes improved annotation, to elucidate epidemiological and phylogenetic relationships.

Material and methods

Strains

The aim was to cover a broad spectrum of M. bovis field strains in the study, including isolates from animals with severe symptoms. Therefore, we selected 36 isolates from different geographical regions of Germany, other countries, and different tissue representing various disease manifestations. All German strains isolated in 2014 and later were collected especially for this study by diagnostic laboratories of several federal states and animal health services and incorporated in the strain collection of the Friedrich-Loeffler Institute, Germany. For comparison, the Cuban strain DL422_88 was included. The basic characteristics of the included strains are given in Table 1. Most isolates were obtained as grown cultures. Further cultivation was done with a modified Hayflick medium containing 20 % horse serum [15] or with commercially available liquid medium containing a phenol-red pH indicator (Mycoplasma Experience Ltd, Bletchingley, UK) at 37 \(^\circ \text {C}\) and 8 % CO\(_{2}\) under static conditions for 2 to 4 days (until color change of the liquid broth to orange or yellow was observed). Only in one case, we obtained a tissue sample from the lung of a calf (16DD0054), which was cultured in liquid broth culture (Mycoplasma Experience Ltd, UK) containing Penicillin G (1000 IU/ml, WDT, Garbsen, Germany) to suppress other bacteria. Liquid cultures were stored in a -80 \(^\circ \text {C}\) freezer and solid agar plates in at 4 \(^\circ \text {C}\).

Table 1 Basic parameters of the M. bovis strains included in this study. CU – Cuba; DE – Germany; PL – Poland; IE – Ireland; BB – Brandenburg; BW – Baden-Württemberg; BY – Bavaria; HE – Hesse; SL – Saarland; MV – Mecklenburg Western Pomerania; NI – Lower Saxony; NRW – Northrhine-Westphalia; RP – Rhineland Palatinate; SN – Saxony; ST – Saxony-Anhalt; TH – Thuringia; SCC – Systemic Circulatory Condition

Description of the disease complex

After calving, the Systemic Circular Condition (SCC) affected cows in medium-sized and large dairy herds. Animals suffered from painful, swollen joints that partially ruptured, edema mainly in the legs and chest area, shortness of breath, and general nervousness while excreting glassy-serous saliva, usually without any fever. Signs of severe mastitis also typically occurred, with udders hardened to a rubbery consistency, milk becoming creamy and sandy and milk yield dropping sharply. Affected udder quarters either became permanently atrophic or their milk production was fully restored after recovery. In the course of the disease, milk cell counts were significantly elevated. Within a few days, some animals could not stand up and finally died from cardiovascular failure. Treatment with antibiotics had no effect while pain-relieving, anti-inflammatory therapy brought some improvement. Feed consumption was only temporarily reduced. Usually, several animals in the herd fell ill one after another with the same symptoms. For example, on a dairy farm with 450 cows, 60 animals showed the same systemic symptoms, of which 20 died. M. bovis was detected in up to 50 % of the diseased cows in a herd. Using commercial ELISA tests, specific antibodies against M. bovis were detected in all diseased animals.

DNA extraction & sequencing

DNA extraction for sequencing with Illumina was done using High Pure PCR Template Preparation Kit from Roche (Mannheim, Germany) according to the manufacturer’s instructions. DNA extraction for Nanopore MinION sequencing was performed using phenol-chloroform extraction as described by Sambrook and Russell [16]. The final DNA extracts were recovered in 50 µl buffer. For the approval of quality and quantity, the Nanodrop spectrophotometer (Thermo Fisher Scientific, Medison, USA) and QUBIT 2.0 fluorometer (Life Technologies Holdings PTE Ltd, Singapore), as well as agarose gel electrophoresis, were used. Illumina sequencing. Genomic DNA (2–10 µg) of each isolate was sent to GATC/Eurofins Genomics (Konstanz, Germany) for genomic library preparation and Illumina MiSeq (2 x 125 bp paired-end or 2 x 150 bp paired-end sequencing) with 5 Million read pairs resulting in an average coverage of around 750x. Raw sequencing data were quality-controlled using FastQC v0.11.8. We used four different MinION flow cells (R9.4.1) numbered 1–4 in Table S1. Nanopore sequencing. Library preparations were done using the 1D genomic DNA by ligation kits (SQK-LSK 108 and SQK-LSK 109) and the native barcoding expansion kits (EXP-NBD103, EXP-NBD104, and EXP-NBD114). In response to an update by ONT, we transitioned to the newer ligation kit SQK-LSK 109 for our nanopore sequencing experiments since the second sequencing run (see Table S1 column F). Briefly, size selection and DNA clean-up were performed using Agencourt AMPure XP beads (Beckman Coulter GmbH, Krefeld, Germany) at a ratio of 1:1 (w:v) before library preparation. Potential nicks in DNA and DNA ends were repaired in a combined step using NEB Next FFPE DNA Repair Mix and NEB Next Ultra II End repair/dA-tailing Module (NewEngland Biolabs, Ipswich, USA) by tripling the incubation time to 15 min at 50 \(^\circ \text {C}\) and 15 min at 65 \(^\circ \text {C}\). The ligation of sequencing adapters followed a subsequent second AMPure bead purification onto prepared ends and a third clean-up step with AMPure beads. Additional barcoding and clean-up steps were performed before adapter ligation. Sequencing buffer and loading beads were added to the library. At the start of sequencing, an initial quality check of the flow cells showed 1571, 1379, 1761, and 1236 active pores. Genomic DNA samples used for loading comprised around 50–150 ng per strain (measured by Qubit 4 Fluorometer; ThermoFisher Scientific, Waltham, USA). The sequencing ran for 48 hrs using the MinKNOW software versions 2.2, 18.12.4, 19.05, 19.10. MinION signals are basecalled and demultiplexed using guppy v6.1.2 model R9.4.1 (only available to ONT customers https://community.nanoporetech.com).

Hybrid de novo assembly pipeline

We developed Mycovista, a pipeline to assemble highly repetitive bacterial genomes such as M. bovis in a hybrid or long-read-only sequencing approach. Our pipeline follows the state-of-the-art de novo genome assembly workflow by incorporating preprocessing of input reads, assembly, polishing, and genome annotation [14]. As input, demultiplexed (nanopore) long reads are required. Illumina paired-end short reads can be used in the hybrid assembly mode. The rapid development of ONT allows now the generation of high-quality assemblies by long read-only data, which is why our pipeline has short-read data as an optional input [17,18,19]. Mycovista returns the assembled genome, a suitable annotation, and general assembly statistics. The pipeline is automated using the workflow management system snakemake v7.3.8 [20] in combination with conda v22.9.0 for reproducible results via stored tool versions in corresponding environment files [21]. All code is publicly available on GitHub (www.github.com/sandraTriebel/mycovista). We ran Mycovista in release version 1.0.

Fig. 1
figure 1

Mycovista – a de novo assembly pipeline. Mycovista can be used in two assembly modes: long and hybrid, requiring only long reads or additionally paired-end short reads, respectively. (0) Basecalled long-read sequencing data (e.g. from ONT) is required. Illumina paired-end short reads can be used as additional input. (1) A raw read quality check (FastQC and Nanoplot) is followed by (2) preprocessing of the input reads (Filtlong, fastp, Trimmomatic) which are then (3) checked regarding their quality. (4) Then, a long-read-only assembly is generated by Flye. (5) Afterwards, the contigs are polished with the preprocessed long reads in several steps (Racon, minimap2, medaka). The assembly can be further postprocessed with short reads. Finally, the (6) quality of the final assembly is assessed (QUAST) followed by a (7) gene annotation (Prokka)

The pipeline consists of seven steps, see Fig. 1: Before using Mycovista, sequencing and basecalling has to be done by the user to provide the input reads. (0) Sequencing and Basecalling: The assembly pipeline can be started in two modes: long requiring only ONT long reads and hybrid where ONT long and Illumina paired-end short reads are needed. MinION signals need to be basecalled and demultiplexed using guppy. We applied guppy v6.1.2 model R9.4.1. using the following parameters: -c dna_r9.4.1_450bps_sup.cfg, --trim_strategy dna, --disable_pings, --disable_qscore_filtering, --calib_detect, --barcode_kits <barcode kits>. (1) Raw Read Quality Check: FastQC v0.11.8 [22] is incorporated for a first quality check of the raw Illumina data. The ONT raw read quality is visualized with Nanoplot v1.41.0 [23]. (2) Preprocessing: Only reads longer than 1,000 nt are considered for the assembly step (filtered via Filtlong v0.2.0 [24]), as M. bovis genomes are known to contain repeats. Illumina raw reads are preprocessed with fastp v0.20.0 [25] for adapter clipping and Trimmomatic v0.39 [26] (sliding window size 4, Phred score quality cut-off of 28, minimum read length 20) to perform quality trimming. However, using Trimmomatic alone is not sufficient to remove all adapter sequences, and thus fastp is included. (3) Preprocessed Read Quality Check: FastQC and Nanoplot are again applied to examine the quality of preprocessed reads and to make a comparison to step (2) possible. (4) Assembly: The filtered long reads are assembled with Flye v2.6 [27]. Optional Illumina reads are not used for assembly but for polishing. (5) Polishing: Post-processing of the long-read assemblies consisted of three steps: (i) four polishing runs with Racon v1.3.2 [28] in combination with minimap2 v2.17 [29] using long reads followed by (ii) one long read-based polishing with medaka v0.11.4 [30] and finally (iii) four polishing runs with Racon and minimap2 using short reads. (6) Assembly Quality Check: Finally, the quality of the so-produced hybrid assemblies is validated by QUAST v5.0.2 [31]. (7) Annotation: All assemblies are annotated using Prokka v1.14.5 [32] with the parameter --gcode 4 to use the required codon table matching characteristics of M. bovis.

Genome synteny analysis

We used Mauve v1.2.0 [33] to detect recombinations in the genome assemblies, such as gene loss, duplication, rearrangement, and horizontal transfer. In order to make our M. bovis genome assembly panel comparable regarding genome rearrangements, we designated the first base of the gene dnaA to be the first base of the genome as described in the study by Mackiewicz et al. [34].

Pangenome analysis and virulence genes

We used PPanGGOLiN v1.2.74 [35] for pangenome analysis of our M. bovis genome assembly panel. PPanGGOLiN generates a core gene set based on a Partitioned Pangenome Graph (PPG), which integrates information about protein-coding genes and their genomic neighborhood. The input genomes are annotated by ppanggolin annotate. Their specific genetic code 4 for Mycoplasmopsis was set with --translation_table 4. Subsequently, clustering was performed using ppanggolin cluster, followed by the graph construction and partitioning (ppanggolin graph, ppanggolin partition). The core genes of each genome were aligned using ppanggolin msa with the parameters --source dna --translation_table 4 and concatenated in one multiple sequence alignment [35], which serves as input for the subsequent reconstruction of the phylogenetic tree. After generating our 36 assemblies, we compared our data with the study by Yair et al. [36], which included 175 assemblies (whole-genome shotgun project PRJNA564939), as well as seven NCBI RefSeq genomes of M. bovis (PG45 NC_014760.1, Hubei-1 NC_015725.1, HB0801 NC_018077.1, CQ-W70 NZ_CP005933.1, 08M NZ_CP019639.1, Ningxia-1 NZ_CP023663.1, JF4278 NZ_LT578453.1) [37,38,39,40,41] and one genome provided by the NCBI (NM2012 CP011348.1).

To compare the gene composition of the 35 European strains concerning phylogenetic clustering, we performed statistical association tests (Fisher’s exact test) for the 35 strains using an R script [42] available in our Mycovista GitHub (www.github.com/sandraTriebel/mycovista/tree/master/scripts/gwas.R). Moreover, we focused on potential virulence factors (such as Vsps) or genes important for the life cycle [10, 43,44,45]. We listed the presence of these genes based on reference-based annotations. Only genes that were assigned with the correct name for the potential virulence factors were taken into account. Multiple occurrences of a gene are possible due to gene duplication or fragmentation. It is noteworthy that Prokka could not annotate vsp genes in our assemblies without a reference genome. This may be due to the absence of those sequence annotations in the database or the very diverse structure of vsp genes.

Phylogenetic tree reconstruction

Phylogenetic trees were reconstructed using IQ-TREE v2.0.3 [46, 47] with a generalized time-reversible (GTR) model based on the alignment of the core gene set determined in pangenome analysis. For this purpose, we included M. agalactiae PG2 (NC_009497.1) as an outgroup [48]. The tree reconstruction of our 36 assemblies was done with 1000 bootstrap runs. For comparison among the 219 M. bovis strains, we reduced the number of bootstraps to 500. Trees and metadata were visualized using iTOL v6.6 [49].

Results

Mycovista in comparison to other tools

In this study, we developed and evaluated a comprehensive de novo genome assembly pipeline to obtain the most accurate and high-quality genomic sequences. The Mycovista pipeline encompassed various stages, including data preprocessing, assembly, polishing, quality assessment, and annotation. To identify the optimal combination of tools and algorithms, we performed a comparison of different state-of-the-art software and methodologies. We assessed the performance of de novo genome assembly tools by analyzing a strain of our dataset (see Table S3). The evaluation criteria included general assembly and annotation statistics. In terms of genome contiguity, Flye [27] performed best with 3 contigs in our benchmark dataset, where the longest contig covered \(\sim\)1 Mb, which is the known genome size of M. bovis. Unicycler [50] and Canu [51] failed to assemble the input reads into a contig covering the expected genome size. While Canu resulted in a small number of contigs, assembly with Unicycler resulted in a more fragmented genome. Thus, we integrated Flye into the Mycovista workflow. For genome annotation, we compared Prokka [32] and Bakta [52]. At the time of the comparison, Bakta annotated twice as many CDSs as Bakta, with many fragmented genes (see Table S4). Prokka, on the other hand, was able to annotate about as many CDS as could be expected for M. bovis. Therefore, we use Prokka for genome annotation. Based on our analyses, we identified the most efficient and reliable tools, which collectively produced highly contiguous and annotated genome assemblies.

Hybrid assembly approach generates contiguous and accurate genomes

Illumina sequencing produced an average of 9,359,418 paired-end reads with a length of 125–150 nt, whereas ONT sequencing produced 317,908 reads on average with a length of about 3,726.8 nt (longest read: 461,771 nt), see Table S1 Illumina & ONT Information.

Based on our hybrid approach, Mycovista achieved high genome contiguity: All 36 hybrid assemblies are comprised of one to maximal three contigs (with the notable exception of 16 contigs in 17DD0020) and have an average N50 of 1,050,036. This is a remarkable improvement given the challenges posed by multiple sequence repeats in the genome [53]. The average genome size of the assemblies in this study was 1.063 Mbp, with the largest genome size being 1.166 Mbp (15DD0218) and the smallest being 0.952 Mbp (15DD0240). Out of 36 M. bovis strains, 23 were assembled into single contigs representing the full chromosome. No contigs of plasmid-origin were found by PlasmidFinder [54], which is in line with other studies of M. bovis [55]. The GC content per assembly ranged from 29.11 % to 29.45 % which is in accordance with previously published M. bovis genomes.

The number of annotated protein-coding genes in our European strain panel varied between 797 and 962, see Table S1, which is in accordance with the expected amount of CDS in M. bovis genomes. We observed varying CDS numbers in strains located at different clades of the phylogenetic tree. In the following, we refer to the assemblies according to their association with a particular clade, see Fig. 3, consistent with the literature [36]. On average, clade 4 shows 50 more CDS than clades 5 and 6 (clade 4: 910, clade 5: 860, clade 6: 856). Almost every genome consists of six rRNA operons (16S, 5S, and 23S rRNA). The number of tRNAs is 34, whereas the annotation displays two more tRNAs with codons for Arg (CCT, TCT) in 15DD0240. Strain 15DD0263 shows four extra tRNAs with codons for Thr (GGT, CGT), Lys (CTT), and Trp (TCA), respectively. Once per Mycoplasmopsis genome, a tmRNA is present, which is representative of other housekeeping ncRNAs illustrating the general quality of the assemblies. Basic statistics about the ONT sequencing, assembly, and annotation results are given in Table S1.

Comparison with global strain panel shows clustering according to geographical origin

To put our assemblies in a global context, we compared the 36 genomes with 175 assemblies from the study by Yair et al. [36] and eight complete assemblies of M. bovis available from the NCBI database (PG45, Hubei-1, HB0801, CQ-W70, NM2012, 08M, Ningxia-1, JF4278). The phylogenetic tree (see Fig. 2) reconstructed from the core gene set (M. agalactiae PG2 included as outgroup) indicates clustering according to geographical origin and is in agreement with the findings of Yair et al. With the exception of the Cuban strain DL422_88 (marked as clade 0), all of our de novo assembled genomes are located in the European cluster (clades 4, 5, and 6), which is consistent with their origin.

For a more detailed analysis of our de novo assembled strain panel, we reconstructed a phylogenetic tree based on the genomes (Fig. 3). The tree reveals the same clustering pattern as observed in the global phylogenetic tree in Fig. 2. Therefore, we will use the same clade designations 0, 4, 5, and 6 when referring to members of our strain panel.

We observed no enrichment of a single clade for isolation source or geographical parameters. Concerning disease manifestations, it is noteworthy that the seven strains associated with SCC were located in two clades: 5.2 and 4. As shown in the right-hand part of Fig. 3, the vsp gene content varies considerably among the M. bovis strains. We used two different sets of annotated variable surface proteins based on strains 8790 (GCF_005061465.1) and PG45 (NC_014760.1), respectively, to search for homologs in our strains (see Fig. S2).

Fig. 2
figure 2

Phylogeny of 219 M. bovis isolates based on the alignment of the core gene set (provided by PPanGGOLiN). We combined our de novo assembled strain panel with 175 assemblies deposited at the whole genome shotgun project PRJNA564939 [36] and eight complete genomes of M. bovis available at the NCBI database. M. agalactiae PG2 was included as an outgroup. The clustering of genomes reflects the geographical origin of strains, with 35 of our own strain panel situated in clades 4, 5, and 6 of the European cluster (cyan), and Cuban strain DL422_88 standing outside. Clade numbers and colors are the same as in the SNP-based phylogenetic tree of the paper by Yair et al. [36]

Fig. 3
figure 3

Phylogenetic tree of our de novo assembled strain panel based on the alignment of the core gene set (provided by PPanGGOLiN). The tree can be divided into three major clusters (4, 5, and 6) and one outlier (Cuban strain DL422_88). We included M. agalactiae PG2 as an outgroup. Biological information such as geographical origin (country, state), source (animal, tissue), pathology, and the presence/absence of vsp genes are shown. Strain designations in boldface denote representative sequences of the clusters used for genome synteny analysis shown in Fig. 4. In the right-hand part, vsp genes in each strain annotated according to NCBI RefSeq strains 8790 and PG45 are depicted. Clade numbers are labeled according to the study by Yair et al. [36]. Bio Information – Biological Information: Central columns show source of isolation (animal, tissue), geographical origin (country, state), and pathology (if available); Co – Cow; DCo – Dead Cow; Ca – Calf; Bu – Bull C – Cattle; BB – Brandenburg; BW – Baden Wurttemberg; BY – Bavaria; HE – Hesse; SL – Saarland; MV – Mecklenburg Western Pomerania; NI – Lower Saxony; NRW – Northrhine-Westphalia; RP – Rhineland Palatinate; SN – Saxony; ST – Saxony-Anhalt; TH – Thuringia; M – Mastitis; P – Pneumonia/Bronchopneumonia; Ab – Abortions; Ar – Arthritis; SCC – Systemic Circulatory Condition

Analysis of the reference-based annotations revealed three general patterns characterizing the occurrence of individual vsp genes in the European strain panel: i) abundant vsps (occurring in more than 30 strains), ii) rare vsps (occurring in less than 5 strains and not clade-specific or absent altogether), as well as iii) those having a preference for one or two clades. Five of the strain 8790-based vsp genes were classified as abundant (WP_233435476, WP_136831431, WP_136831844, WP_136832078, WP_136832058), four as rare (WP_211235702, WP_136832641, WP_136832633, WP_233435459), and the following genes showed a clade preference: WP_233435477 (for clades 5 and 6), WP_136831 985 (4 and 6), WP_136832637 (4), WP_136832228 (6), and WP_136831846 (5). When looking for vsp homologs of strain PG45, vsps G, K, and L were present in more than 30 genomes in our strain panel and in all clades, respectively. Members of subclade 6.3 were found to harbor several vspA gene copies in their genome. The vspB gene was identified only in one strain, while vspF was absent in all 36 strains. Among vsps with clade preference, vspE and vspM genes seemed to be confined to clade 4, whereas vspJ and vspO were only encountered in clade 5. Total count of genes are shown in Table S2.

Comparison of genome synteny among the European strain panel

Analysis of the global organization of the 36 de novo assembled genomes revealed rearrangements when comparing clusters or subclusters. Notably, there are characteristic features that are consistent within the genomes of a cluster, which also indicates that they are not the result of assembly errors. To illustrate clade-specific changes in genome synteny, we selected eight strains representing the major clades for calculating an alignment using Mauve, see Fig S2.

While the first three blocks 5’ and the last two blocks 3’ are highly conserved among all strains in terms of synteny, the central genomic region, which contains more than 10 locally collinear blocks, displays the highest divergence. As expected, the Cuban strain DL422_88 is distinct from all other European isolates and has a markedly different synteny compared to the rest of the strain panel. A major characteristic of clade 6 strains (15DD0210, 15DD0261, 15DD0240) is the inverted central part of the genome, in which the blocks appear in the reverse complement orientation relative to the other genomes. This region displays the highest degree of divergence among clade members in comparison to clades 4 and 5 (Fig. S2). Strain 15DD0160 (and two other strains from clade 4, see Fig. S2) also appears to have an inverted central block, similar to the strains of clade 6. In contrast, the remaining representatives of clades 4 (15DD0234 and 16DD0186) and 5 (15DD0165) share the central genomic blocks arranged in the forward orientation, see Fig. 4. There is also variation among individual strains within the clade (see Fig. S2). The genomes of clade 4 are 50,000 to 100,000 bp larger than those of clades 5 and 6, which explains the higher number of CDSs in clade 4 strains. These genomes contain a number of segments depicted as blank spaces that have no homologs in the other strains (Fig. 4 and S2).

Fig. 4
figure 4

Multiple genome alignment of eight strains representing the major clades using Mauve. Each genome was linearized and normalized, with homologous segments (locally collinear blocks) shown as colored rectangles. Inverted regions are set below those that match the neighboring genome in the forward orientation. Lines collate aligned segments between genomes. The alignment of all 36 assemblies done in this study is shown in Fig. S2

Fig. 5
figure 5

Gene association analysis (Fisher’s exact test) of our European assembly panel in correlation to phylogeny. We detected 108 genes to be significantly associated with at least one cluster. Of these genes, 25 are linked to a gene product or at least a class of products, including various enzymes, e.g. proteases and nucleases, as well as ten variable surface lipoproteins and other surface proteins. Some genes are listed more than once when truncated versions are also associated with a cluster

Pangenome analysis reveals cluster-associated genes including virulence factors

We identified a pangenome of our 36 assemblies consisting of 1,143 gene families, composed of 598 core genes (i.e., genes present in all 36 genomes), 105 shell genes present in \(\sim\) 55 % genomes, and 351 cloud genes present at low frequency (\(\sim\) 8 %). Gene association analysis revealed 108 to be significantly associated with at least one cluster (4, 5, 6), i.e. are either significantly enriched or depleted in the strains of the analyzed cluster (see Fig. S3). Of those genes, 25 could be linked to a gene product or at least a class of products. These clade-specific genes include various enzymes, e.g. proteases and nucleases, as well as ten variable surface lipoproteins and other surface proteins (see Fig. 5). In particular, some vsp genes show correlations to phylogenetic clusters. In Fig. 3, we can observe vsp WP_136832228 to be only present in clade 6 and Cuban strain DL422_88 (clade 0). The genes vspJ and vspO are only present in clade 5. Genes vsp WP_136832637 and vspE are associated with clade 4.

As shown in Fig. 5, genes encoding iron-sulfur cluster carrier protein, tyrosine recombinase XerD, conjugal transfer protein TraE, ATP-dependent zinc metalloprotease FtsH, transglutaminase domain-containing protein, and DNA cytosine methyltransferase seem to be only annotated in genomes of clade 4. Gene xerC coding for tyrosine recombinase XerC is annotated three to four times in clades 5 and 6, but only two times in clade 4. We observed these differences in the annotation files as well as in the gene association analysis in which one annotation of xerC was detected as absent in clade 4. The gene vsp WP_136831846 is absent in clade 6. Genes vspN were annotated multiple times in some genomes. One of them occurred only in clade 5 strains, and another one was only present in clade 4. We observed similar results for vsp WP_136831985, see Fig. 5, which was never annotated in group 5 genomes, but found once in each of clades 4 and 6, respectively. The gene vsp WP_136831431 was annotated twice in clades 4 and 5 and in Cuban strain DL422_88 (clade 0), but only once in clade 6 as revealed by Fisher’s exact test.

Discussion

De novo assembly

Mycovista assembles bacterial genomes using a hybrid approach to tackle the problem of resolving highly repetitive regions. The use of long reads (ONT used here; in theory PacBio too, but not tested) in the initial assembly is beneficial for generating contiguous genomes, and the polishing steps involving short reads improve the sequence accuracy of the genomes. This aligns with state-of-the-art studies and current best practices in assembling high-quality bacterial genomes [14]. General assembly statistics and genome annotation are additional steps in the pipeline for further analysis. With our method, we were able to assemble 36 M. bovis strains examined in this study into contiguous genomes and to ensure high sequence accuracy through postprocessing of the assemblies, resulting in suitable annotations.

Phylogenetic clades

Comparison in a global context showed clustering according to geographical origin, with the newly assembled genomes positioned in the European cluster (comprising clades 4, 5, and 6) of the tree. In contrast, the genome of the Cuban strain DL422_88 proved phylogenetically more distant and did not cluster with any other genome.

The geographic clustering, however, is not of high resolution, i.e. not country-specific, which is probably due to the extensive international animal trade throughout Europe. The fact that the European cluster also includes strains from the US (e.g. type strain PG45) could be seen as an indication that phylogenetic processes in M. bovis occur at a slow pace.

Analysis of our assemblies also revealed cluster association of several genes, including the vsps, whose products are presumed virulence factors. Other correlations between phylogeny and strain properties were not observed. Genomic data supporting tissue specificity of strains or association with specific disease manifestations have not been provided in the literature so far. The identification of a mastitis-dominant lineage of M. bovis strains by Yair et al. [36] was certainly due to the relatively small size of the cattle population in Israel, where calves were imported from five to seven countries only. Nevertheless, strains associated with mastitis from that study did not cluster on a single clade, which indicates that genomic factors alone are probably insufficient to characterize a strain’s association with a particular clinical manifestation.

Genome, core genome, pangenome

Genomes of Mycoplasmopsis spp. are the smallest among bacteria. With an average length of 1.063 Mbp, our panel of 36 strains has genomes of the expected size, with a core genome of 598 genes and a pangenome of 1,143 (Table S1). The average number of annotated protein-coding genes of 880 (Table S1) is higher than in other studies, probably due to the hybrid assembly approach, which allowed complete genome assembly in most cases and improved annotation. Kumar et al. analyzed a large set of 250 M. bovis strains, mainly from the US and Australia, appeared to be phylogenetically more heterogeneous than our European panel. These authors used Illumina sequence data and reported an average CDS number of 770, core and pangenome of 283 and 1,186, respectively [56]. The data of our own study comprising 219 strains are comparable: 327 core genes and 1,623 genes in the pangenome (the latter due to annotation, see above).

Essential vs. dispensable genes

From a formal point of view, the core genome elements are considered indispensable for the organism’s survival [57]. From a biological viewpoint, genes encoding replication and translation factors, as well as elements of metabolic pathways, are considered essential because they belong to the fundamental cellular machinery of bacteria. Josi et al. [58] identified 352 out of 900 M. bovis genes as essential, most of them involved in nucleotide metabolism or biosynthesis of secondary metabolites and amino acids. Using this definition, the clade-specific genes identified in our study (partly given in Fig. 5) would mainly be classified as non-essential.

vsp locus

No precise figures on the varying size of the entire vsp locus in M. bovis genomes are available. Early data from Lysnyansky et al. [7, 8] indicate that the size of all 13 sequences coding for Vsp polypeptide chains in strain PG45 is about 11 kbp, to which the promoter (150 bp each) and signal peptide (75 bp) regions have to be added (total approx. 3 kbp). This means that 1.4 % of the genome would be used to encode Vsp family members. The proportion could be higher in strains equipped with multiple vsp gene copies and lower in strains lacking individual vsp genes.

Josi et al. classified vsp genes as non-essential for the organism [58]. This may be due to the software used in their study, which could not assign reads from highly repetitive regions to a specific position in the genome, and, therefore, did not consider vsps. From the biological perspective, it seems more likely that the vsp locus is not dispensable since individual members were suggested to play a role in cytoadhesion and evasion of the host immune response [10, 59]. The fact that M. bovis strains can swiftly alter their Vsp repertoire and protein chain length in the face of host or environmental challenges could also indicate that the underlying genes are essential for the organism.

When Kumar et al. used the sequences of the 13 vsp genes of strain PG45 as BLAST queries they found that none of their strains harbored the complete vsp gene set [56]. While it is known that the number of vsp family members can be reduced in certain M. bovis strains the high degree of sequence variation in the vsp genes may have contributed to the low number of hits in that study.

SCC as a largely unknown manifestation of M. bovis infection

The symptomatology observed in SCC has not been described before and represents an additional clinical picture that might be associated with M. bovis infection. The signs of this new condition primarily include severe edema, particularly in the thoracic and abdominal regions, as well as arthritis, even in adult animals (see Fig. S1). The formation of edema indicates involvement of the cardiovascular system. We observed several fatalities in dairy cows with SCC in conjunction with M. bovis isolation. Our hypothesis was that M. bovis strains isolated from animals presenting with SCC carry genomic traits responsible for higher virulence. However, using our comparative genomic approach involving genome-wide association analysis, such traits could not be identified. Therefore, the question of whether M. bovis contributes directly or indirectly to SCC pathogenesis remains open. In addition to SCC, several cases have been reported, in which M. bovis was not only implicated in Bovine Respiratory Disease (BRD) in calves but also caused fibrinous bronchopneumonia in adult animals reminiscent of Contagious Bovine Pleuropneumonia (CBPP) [5, 6]. This underlines the importance of M. bovis as a differential diagnosis for the reportable CBPP, which should be conducted by means of molecular identification tools.

Availability of data and materials

Mycovista is available on GitHub https://github.com/sandraTriebel/mycovista. Our de novo assembly panel (36 assemblies) is provided in the NCBI BioProject PRJNA954308.

References

  1. Nicholas RAJ, Ayling RD. Mycoplasma bovis: disease, diagnosis, and control. Res Vet Sci. 2003;74(2):105–12. https://doi.org/10.1016/S0034-5288(02)00155-8.

  2. Pfützner H, Sachse K. Mycoplasma bovis as an agent of mastitis, pneumonia, arthritis and genital disorders in cattle. Rev Sci Tech. 1996;15(4):1477–94.

    Article  PubMed  Google Scholar 

  3. Nicholas R, Ayling R, McAuliffe L. Bovine respiratory disease in Mycoplasma diseases of ruminants. CABI Wallingford; 2008.

  4. Maunsell FP, Woolums AR, Francoz D, Rosenbusch RF, Step DL, Wilson DJ, et al. Mycoplasma bovis Infections in Cattle. J Vet Intern Med. 2011;25(4):772–83. https://doi.org/10.1111/j.1939-1676.2011.0750.x.

  5. Heller M, Kammerer R, Sehl J, Teifke JP, Schubert E. Annual Report of the National Reference Laboratory for CBPP. Friedrich-Loeffler-Institut; 2017.

  6. Heller M, Schubert E, Schnee C. Annual Report of the National Reference Laboratory for CBPP. Friedrich-Loeffler-Institut; 2019.

  7. Lysnyansky I, Ron Y, Yogev D. Juxtaposition of an Active Promoter to vsp Genes via Site-Specific DNA Inversions Generates Antigenic Variation in Mycoplasma bovis. J Bacteriol. 2001;183(19):5698–708. https://doi.org/10.1128/JB.183.19.5698-5708.2001.

  8. Lysnyansky I, Sachse K, Rosenbusch R, Levisohn S, Yogev D. The vsp Locus of Mycoplasma bovis: Gene Organization and Structural Features. J Bacteriol. 1999;181(18):5734–5741. https://doi.org/10.1128/JB.181.18.5734-5741.1999.

  9. Nussbaum S, Lysnyansky I, Sachse K, Levisohn S, Yogev D. Extended Repertoire of Genes Encoding Variable Surface Lipoproteins in Mycoplasma bovis Strains. Infect Immun. 2002;70(4):2220–5. https://doi.org/10.1128/IAI.70.4.2220-2225.2002.

  10. Bürki S, Frey J, Pilo P. Virulence, persistence and dissemination of Mycoplasma bovis. Vet Microbiol. 2015;179(1):15–22. Special Issue: VETPATH 2014 - Pathogenesis of Bacterial Infections of Animals. https://doi.org/10.1016/j.vetmic.2015.02.024.

  11. Calcutt MJ, Lysnyansky I, Sachse K, Fox LK, Nicholas RAJ, Ayling RD. Gap analysis of Mycoplasma bovis disease, diagnosis and control: An aid to identify future development requirements. Transboundary Emerg Dis. 2018;65(S1):91–109. https://doi.org/10.1111/tbed.12860.

  12. Jain M, Olsen HE, Paten B, Akeson M. The Oxford Nanopore MinION: delivery of nanopore sequencing to the genomics community. Genome Biol. 2016;17(1):239.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Loman NJ, Quick J, Simpson JT. A complete bacterial genome assembled de novo using only nanopore sequencing data. Nat Methods. 2015;12(8):733–5.

    Article  CAS  PubMed  Google Scholar 

  14. Wick RR, Judd LM, Holt KE. Assembling the perfect bacterial genome using Oxford Nanopore and Illumina sequencing. PLoS Comput Biol. 2023;19(3): e1010905. https://doi.org/10.1371/journal.pcbi.1010905.

  15. Freundt EA. Culture media for Classic Mycoplasmas. In: Methods in Mycoplasmology. Elsevier; 1983. p. 127–135. https://doi.org/10.1016/B978-0-12-583801-6.50029-9.

  16. Sambrook J, Russell DW. Molecular cloning: a laboratory manual, vol. 1. 3rd ed. New York: Cold Spring Harbor Laboratory Press; 2001.

  17. Wick RR, Judd LM, Gorrie CL, Holt KE. Completing bacterial genome assemblies with multiplex MinION sequencing. Microb Genomics. 2017;3(10). https://doi.org/10.1099/mgen.0.000132.

  18. Sereika M, Kirkegaard RH, Karst SM, Michaelsen TY, Sørensen EA, Wollenberg RD, et al. Oxford Nanopore R10.4 long-read sequencing enables the generation of near-finished bacterial genomes from pure cultures and metagenomes without short-read or reference polishing. Nat Methods. 2022;19(7):823–826. https://doi.org/10.1038/s41592-022-01539-7.

  19. Vereecke N, Bokma J, Haesebrouck F, Nauwynck H, Boyen F, Pardon B, et al. High quality genome assemblies of Mycoplasma bovis using a taxon-specific Bonito basecaller for MinION and Flongle long-read nanopore sequencing. BMC Bioinformatics. 2020;21(1):517. https://doi.org/10.1186/s12859-020-03856-0.

  20. Köster J, Rahmann S. Snakemake—a scalable bioinformatics workflow engine. Bioinformatics. 2012;28(19):2520–2. https://doi.org/10.1093/bioinformatics/bts480.

  21. Anaconda Software Distribution [Internet]. Anaconda Documentation. Anaconda Inc. 2020. Available from https://docs.anaconda.com/. Version 22.9.0 released on 14.09.2022.

  22. Andrews S, Krueger F, Segonds-Pichon A, Biggins L, Krueger C, Wingett S. FastQC [Internet]. Babraham: Babraham Institute; 2012. Available from https://www.bioinformatics.babraham.ac.uk/projects/fastqc/ and conda. Version 0.11.8 released on 04.10.2018.

  23. De Coster W, D’Hert S, Schultz DT, Cruts M, Van Broeckhoven C. NanoPack: visualizing and processing long-read sequencing data. Bioinformatics. 2018;34(15):2666–9. https://doi.org/10.1093/bioinformatics/bty149.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Wick R. Filtlong [Internet]. 2018. Available from https://github.com/rrwick/Filtlong and conda. Version 0.2.0 released on 4.01.2018.

  25. Chen S, Zhou Y, Chen Y, Gu J. fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics. 2018;34(17):i884–90. https://doi.org/10.1093/bioinformatics/bty560.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30(15):2114–20. https://doi.org/10.1093/bioinformatics/btu170.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Kolmogorov M, Yuan J, Lin Y, Pevzner PA. Assembly of long, error-prone reads using repeat graphs. Nat Biotechnol. 2019;37(5):540–6.

    Article  CAS  PubMed  Google Scholar 

  28. Vaser R, Sović I, N N, Šikić M. Fast and accurate de novo genome assembly from long uncorrected reads. Genome Res. 2017;27:737–46.

  29. Li H. Minimap2: pairwise alignment for nucleotide sequences. Bioinformatics. 2018;34(18):3094–100. https://doi.org/10.1093/bioinformatics/bty191.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Oxford Nanopore Technologies Ltd. medaka. 2018. Available from https://github.com/nanoporetech/medaka and conda. Version 0.11.4 released on 14.01.2020.

  31. Gurevich A, Saveliev V, Vyahhi N, Tesler G. QUAST: quality assessment tool for genome assemblies. Bioinformatics. 2013;29(8):1072–5. https://doi.org/10.1093/bioinformatics/btt086.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Seemann T. Prokka: rapid prokaryotic genome annotation. Bioinformatics. 2014;30(14):2068–9. https://doi.org/10.1093/bioinformatics/btu153.

    Article  CAS  PubMed  Google Scholar 

  33. Darling ACE, Mau B, Blattner FR, Perna NT. Mauve: multiple alignment of conserved genomic sequence with rearrangements. Genome Res. 2004;14(7):1394–403.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Mackiewicz P, Zakrzewska-Czerwińska J, Zawilak A, Dudek MR, Cebrat S. Where does bacterial replication start? Rules for predicting the oriC region. Nucleic Acids Res. 2004;32(13):3781–91. https://doi.org/10.1093/nar/gkh699.

  35. Gautreau G, Bazin A, Gachet M, Planel R, Burlot L, Dubois M, et al. PPanGGOLiN: Depicting microbial diversity via a partitioned pangenome graph. PLoS Comput Biol. 2020;16(3):1–27. https://doi.org/10.1371/journal.pcbi.1007732.

    Article  CAS  Google Scholar 

  36. Yair Y, Borovok I, Mikula I, Falk R, Fox LK, Gophna U, et al. Genomics-based epidemiology of bovine Mycoplasma bovis strains in Israel. BMC Genomics. 2020;21(1):70. https://doi.org/10.1186/s12864-020-6460-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Wise KS, Calcutt MJ, Foecking MF, Röske K, Madupu R, Methé BA. Complete Genome Sequence of Mycoplasma bovis Type Strain PG45 (ATCC 25523). Infect Immun. 2011;79(2):982–3.

  38. Li Y, Zheng H, Liu Y, Jiang Y, Xin J, Chen W, et al. The Complete Genome Sequence of Mycoplasma bovis Strain Hubei-1. PLoS ONE. 2011;6(6):1–10. https://doi.org/10.1371/journal.pone.0020999.

    Article  CAS  Google Scholar 

  39. Qi J, Guo A, Cui P, Chen Y, Mustafa R, Ba X, et al. Comparative Geno-Plasticity Analysis of Mycoplasma bovis HB0801 (Chinese Isolate). PLoS ONE. 2012;7(5):1–13. https://doi.org/10.1371/journal.pone.0038239.

    Article  CAS  Google Scholar 

  40. Chen S, Hao H, Zhao P, Gao P, He Y, Ji W, et al. Complete Genome Sequence of Mycoplasma bovis Strain 08M. Genome Announc. 2017;5(19):e00324–17.

  41. Sun P, Luo H, Zhang X, Xu J, Guo Y, He S. Whole-Genome Sequence of Mycoplasma bovis Strain Ningxia-1. Genome Announc. 2018;6(4):e01367–17.

  42. R Core Team. R: A Language and Environment for Statistical Computing. Vienna: R Foundation for Statistical Computing; 2021.

  43. Rasheed MA, Qi J, Zhu X, Chenfei H, Menghwar H, Khan FA, et al. Comparative Genomics of Mycoplasma bovis Strains Reveals That Decreased Virulence with Increasing Passages Might Correlate with Potential Virulence-Related Factors. Front Cell Infect Microbiol. 2017;7. https://doi.org/10.3389/fcimb.2017.00177.

  44. Parker AM, Shukla A, House JK, Hazelton MS, Bosward KL, Kokotovic B, et al. Genetic characterization of Australian Mycoplasma bovis isolates through whole genome sequencing analysis. Vet Microbiol. 2016;196:118–25. https://doi.org/10.1016/j.vetmic.2016.10.010.

  45. Behrens A, Heller M, Kirchhoff H, Yogev D, Rosengarten R. A family of phase- and size-variant membrane surface lipoprotein antigens (Vsps) of Mycoplasma bovis. Infect Immun. 1994;62(11):5075–84. https://doi.org/10.1128/iai.62.11.5075-5084.1994.

  46. Nguyen LT, Schmidt HA, von Haeseler A, Minh BQ. IQ-TREE: A Fast and Effective Stochastic Algorithm for Estimating Maximum-Likelihood Phylogenies. Mol Biol Evol. 2014;32(1):268–74. https://doi.org/10.1093/molbev/msu300.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Hoang DT, Chernomor O, von Haeseler A, Minh BQ, Vinh LS. UFBoot2: Improving the Ultrafast Bootstrap Approximation. Mol Biol Evol. 2017;35(2):518–22. https://doi.org/10.1093/molbev/msx281.

    Article  CAS  PubMed Central  Google Scholar 

  48. Sirand-Pugnet P, Lartigue C, Marenda M, Jacob D, Barré A, Barbe V, et al. Being Pathogenic, Plastic, and Sexual while Living with a Nearly Minimal Bacterial Genome. PLoS Genet. 2007;3(5):1–15. https://doi.org/10.1371/journal.pgen.0030075.

  49. Letunic I, Bork P. Interactive Tree Of Life (iTOL): an online tool for phylogenetic tree display and annotation. Bioinformatics. 2006;23(1):127–8. https://doi.org/10.1093/bioinformatics/btl529.

    Article  CAS  PubMed  Google Scholar 

  50. Wick RR, Judd LM, Gorrie CL, Holt KE. Unicycler: Resolving bacterial genome assemblies from short and long sequencing reads. PLoS Comput Biol. 2017;13(6):e1005595.

  51. Koren S, Walenz BP, Berlin K, Miller JR, Bergman NH, Phillippy AM. Canu: scalable and accurate long-read assembly via adaptive k -mer weighting and repeat separation. Genome Res. 2017;27(5):722–36. https://doi.org/10.1101/gr.215087.116.

  52. Schwengers O, Jelonek L, Dieckmann MA, Beyvers S, Blom J, Goesmann A. Bakta: rapid and standardized annotation of bacterial genomes via alignment-free sequence identification. Microb Genomics. 2021;7(11): 000685. https://doi.org/10.1099/mgen.0.000685.

    Article  CAS  Google Scholar 

  53. Schmid M, Frei D, Patrignani A, Schlapbach R, Frey JE, Remus-Emsermann MNP, et al. Pushing the limits of de novo genome assembly for complex prokaryotic genomes harboring very long, near identical repeats. Nucleic Acids Res. 2018;46(17):8953–65. https://doi.org/10.1093/nar/gky726.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Carattoli A, Hasman H. In: de la Cruz F, editor. PlasmidFinder and In Silico pMLST: Identification and Typing of Plasmid Replicons in Whole-Genome Sequencing (WGS). New York: Springer US; 2020. p. 285–294. https://doi.org/10.1007/978-1-4939-9877-7_20.

  55. Breton M, Tardy F, Dordet-Frisoni E, Sagne E, Mick V, Renaudin J, et al. Distribution and diversity of mycoplasma plasmids: lessons from cryptic genetic elements. BMC Microbiology. 2012;12(1):257. https://doi.org/10.1186/1471-2180-12-257.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Kumar R, Register K, Christopher-Hennings J, Moroni P, Gioia G, Garcia-Fernandez N, et al. Population Genomic Analysis of Mycoplasma bovis Elucidates Geographical Variations and Genes associated with Host-Types. Microorganisms. 2020;8(10). https://doi.org/10.3390/microorganisms8101561.

  57. Segerman B. The genetic integrity of bacterial species: the core genome and the accessory genome, two different stories. Front Cell Infect Microbiol. 2012;2. https://doi.org/10.3389/fcimb.2012.00116.

  58. Josi C, Bürki S, Vidal S, Dordet-Frisoni E, Citti C, Falquet L, et al. Large-Scale Analysis of the Mycoplasma bovis Genome Identified Non-essential, Adhesion- and Virulence-Related Genes. Front Microbiol. 2019;10. https://doi.org/10.3389/fmicb.2019.02085.

  59. Sachse K, Helbig JH, Lysnyansky I, Grajetzki C, Müller W, Jacobs E, et al. Epitope Mapping of Immunogenic and Adhesive Structures in Repetitive Domains of Mycoplasma bovis Variable Surface Lipoproteins. Infect Immun. 2000;68(2):680–7. https://doi.org/10.1128/IAI.68.2.680-687.2000.

Download references

Acknowledgements

We thank all the colleagues mentioned in Table 1, who kindly provided strains for this study. Susann Bahrmann is acknowledged for excellent technical assistance.

Funding

Open Access funding enabled and organized by Projekt DEAL. This work was funded by the Balance of the Microverse (EXC 2051, 390713860), NFDI4microbiota (NFDI 28/1, 460129525), iDiv (FZT 118/2, 202548816), Carl-Zeiss-Stiftung (FKZ 0563-2.8/378/2), and DigLeben (5575/10-9).

Author information

Authors and Affiliations

Authors

Contributions

KS, MHö, and MM designed the study. KS did conventional sequence assemblies at the initial stage. MHö and ST developed the new assembly pipeline. ST conducted the assembly and processing of raw sequence data. MHe and CS provided the strains and submitted them for sequencing. CD did Nanopore sequencing and wetlab work. MW conducted pairwise association analysis. ST, KS, MHö, and MM wrote the manuscript. All authors read and approved the final version of the manuscript.

Corresponding author

Correspondence to Manja Marz.

Ethics declarations

Ethics approval and consent to participate

Samples were collected by attending veterinarians as part of diagnostic investigations and as no additional pain, suffering or harm was inflicted on the animals, no approval from an ethics committee was required under national law. Written consent was obtained from the animal owners.

Consent for publication

Not applicable.

Competing interests

MHö is a co-founder of nanozoo GmbH and holds shares in the company. The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Triebel, S., Sachse, K., Weber, M. et al. De novo genome assembly resolving repetitive structures enables genomic analysis of 35 European Mycoplasmopsis bovis strains. BMC Genomics 24, 548 (2023). https://doi.org/10.1186/s12864-023-09618-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-023-09618-5

Keywords