Genomic characterization of eight novel Bartonella species from bats and ectoparasites reveals phylogenetic diversity and host adaptation

0
8

?
This is an uncorrected proof.
Figures
Abstract
Bats are recognized as hosts of diverse Bartonella species, and Bartonella are known agents of human disease. Here, we reported the isolation of 26 Bartonella strains belonging to eight distinct species from two bat species (Myotis laniger and M. adversus) and their ectoparasites (bat flies and bat mites) collected from Chongming County, Hubei Province, China. We obtained whole genome sequences of the bat-borne Bartonella spp. with second and third generation sequencing. The genome size of the bat-borne Bartonella spp. ranged from 1.39 to 1.98 Mb, with the GC content from 35.30%-38.92%. Phylogenetic analysis revealed that the bat-borne Bartonella spp. were divergent from currently known Bartonella spp., and the Average Nucleotide Identity (ANI) values were all below 95%, indicating that they were all novel species. Comparative genomic analysis underscored distinct characteristics of bat-borne Bartonella spp. in cell motility, replication, recombination, and the biogenesis of the cell wall/membrane/envelope, suggesting that bat-borne Bartonella spp. exhibited unique host adaptability. Notably, the virulence factor genes of two bat-borne Bartonella spp., strains B23 and 39, were highly analogous to those of the currently known Bartonella henselae and Bartonella quintana, indicating that they might be pathogenic to humans and animals. Altogether, our findings significantly broaden the diversity of the Bartonella genus, and provide new insights into the host specificity and evolutionary relationship of bat-borne Bartonella spp..
Author summary
Bats and their ectoparasites are natural hosts of many bacteria that may cause disease in humans, especially in tropical and subtropical regions. In this study, we discovered eight previously unknown Bartonella species from bats and their parasites in China. By analyzing their genomes, we found that these bacteria are genetically distinct and carry features suggesting they have adapted to specific hosts. Some also contain genes similar to those found in known human pathogens, raising concern about their potential to infect humans or animals. This research improves our understanding of bacterial diversity in wildlife and highlights the importance of monitoring lesser-known microbes that may pose future health risks.
Citation: Wang W-J, Gao Q, Guo B-Z, Xiao X, Han H-J (2025) Genomic characterization of eight novel Bartonella species from bats and ectoparasites reveals phylogenetic diversity and host adaptation. PLoS Negl Trop Dis 19(10): e0013646. https://doi.org/10.1371/journal.pntd.0013646
Editor: Vladimir L. Motin, University of Texas Medical Branch, UNITED STATES OF AMERICA
Received: July 5, 2025; Accepted: October 15, 2025; Published: October 23, 2025
Copyright: © 2025 Wang et al. This is an open access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Data Availability: All genome sequences generated in this study have been deposited in GenBank. All accession numbers have been provided in the main text and in S3 Table. The raw PCR positivity data for individual bats and ectoparasites are provided in S1 File.
Funding: This project was supported by the Shandong Provincial Natural Science Foundation (ZR2023QH318 to HH) and the Hubei Provincial Natural Science Foundation Collaborative Fund Project (No. 2023AFD164 to XX). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
Competing interests: The authors have declared that no competing interests exist.
Introduction
Bartonella spp. are a group of Gram-negative, facultative intracellular bacteria, which infect and replicate in endothelial cells and erythrocytes [1]. Currently, the genus Bartonella comprises 59 confirmed species and 17 provisional species and many Bartonella species are associated with human diseases including Bartonella bacilliformis (Carrion′s disease), B. henselae (cat-scratch disease), and B. quintana (trench fever) [2,3]. Moreover, emerging novel Bartonella spp. are reported to be responsible for blood-culture-negative endocarditis [4,5]. Bartonella have a wide range of mammalian hosts, including rodents, cats, and dogs [6–8]. Bats, belonging to the Order Chiroptera, are important reservoirs of emerging infectious diseases [9]. Globally, a diversity of Bartonella spp. are identified in bats and their ectoparasites, highlighting the role of bats as hosts of Bartonella species [10,11]. Currently, our understanding of bat-borne Bartonella spp. is limited due to the scarcity of whole genome data. Such genomic data are important for comprehensive characterization of their genetic diversity, the identification of key virulence determinants, and the elucidation of evolutionary adaptations underlying host specificity and potential zoonotic transmission. In our previous work, a diversity of novel Bartonella spp. in bats and their ectoparasites in China was identified using multi-locus PCR approaches [12,13]. To better understand bat-borne Bartonella, we herein aim to isolate Bartonella from bats and their ectoparasites, and to obtain the whole genome sequences of bat-borne novel Bartonella spp. for phylogenetic analysis and comparative genomic analysis. Our findings will significantly broaden the diversity of the Bartonella genus and provide new insights into the host specificity and evolutionary relationships of bat-borne Bartonella spp..
Materials and methods
Ethics statement
The collection of bats for Bartonella isolation was approved by the Ethics Committee of Shandong First Medical University (W202303060236), and all efforts were made to minimize discomfort to bats. The study was conducted in strict accordance with the institutional guidelines.
Sampling and species identification of bats and their ectoparasites
Bats were sampled from Chongming County, Hubei Province, China, using mist nets placed at the entrances of karst caves during peak activity periods of bats (at sunset and before dawn). Nets were continuously monitored, and captured bats were promptly removed upon capture. Captured bats were placed in cloth bags, euthanized by inhalation of ethyl ether in the field with efforts to minimize animal discomfort, and transported on ice to the laboratory for further analysis. Bat flies were collected from the fur of bats, and bat mites were obtained from their wing membranes. Bat spleens were collected via necropsy and specimens were stored at −80°C until use. Bats and their ectoparasites were identified by PCR amplification and subsequent sequencing of the cytochrome B (cytB) and cytochrome oxidase subunit I (COI) genes, respectively [13].
Bartonella isolation and molecular identification
As an ongoing program aimed at identifying pathogens in bats, the spleen was used for Bartonella isolation in this study, which typically harbors higher bacterial loads than blood, facilitating the detection and isolation of Bartonella [14,15]. Bat spleen samples were ground with 200 μL Schneider’s insect medium (Sigma-Aldrich, Germany) supplemented with 10% fetal bovine serum (BBI, China) and 2% amphotericin B (Solarbio, China). Bat ectoparasites were washed with 75% alcohol and then washed twice with PBS before being homogenized using a handheld electric homogenizer. An aliquot of 100 μL homogenate was inoculated onto a Columbia blood agar plate (Oxoid, UK) supplemented with 10% defibrinated sheep blood (Solarbio, China). The plates were cultured in a humidified incubator at 37°C, and 5% CO2, and were checked daily for Bartonella colonies for a month. Each suspected Bartonella isolate was obtained from a single colony, subcultured through three passages, and identified by PCR amplification and sequencing of the 16S rRNA, gltA, rpoB, and ftsZ genes with primers as described previously [13,16]. Based on the primer schemes employed in previous studies, modifications were made in this study. The primers used for 16S rRNA, gltA, rpoB, and ftsZ are provided in S6 Table. Maximum likelihood phylogenetic trees were constructed for the 16S rRNA, gltA, rpoB, and ftsZ genes using IQ-TREE with 1,000 bootstrap replicates [17]. The best-fitting nucleotide substitution model for each gene was selected based on the Bayesian Information Criterion (BIC), resulting in K2P+I + G4, TIM3 + F + R3, TN + F + R3, and HKY + F + R2 for the 16S rRNA, gltA, rpoB, and ftsZ genes, respectively.
Whole-genome sequencing and phylogenetic analysis
Isolated Bartonella spp. were scraped off the Columbia blood agar plates, and genomic DNA was extracted with NucleoBond HMW DNA Kit (Takara, Japan). Genomic DNA was subjected to both short- and long-read sequencing. For short-read sequencing, DNA was randomly fragmented to approximately 300 bp to construct paired-end libraries, which were sequenced on the MGI DNBSEQ-T7 platform using a 150 bp paired-end strategy. For long-read sequencing, DNA was sheared with g-TUBEs (Covaris, USA), size-selected and purified with AMPure PB beads, and libraries were prepared using the SMRTbell Express Template Prep Kit 2.0 (Pacific Biosciences, USA). The polymerase–primer complexes were generated with the Sequel II Binding Kit 2.0, and sequencing was performed on the PacBio Sequel II platform according to the manufacturer’s instructions. Fastp was used to remove adapter sequences and low-quality sequences from the second-generation data, and Fastplong was used for quality control on the third-generation sequencing data [18,19]. The genome was assembled using Spades [20], and the completeness and accuracy of genome assembly were assessed using BUSCO and CheckM, respectively [21,22]. Prokka was used for protein annotation [23]. Orthofinder was employed to identify single-copy orthologous proteins, each orthologous gene was aligned using Mafft with the FFT-NS-2 algorithm and the same parameters across all genes [24,25]. The resulting alignments were concatenated into a supermatrix, which was then used to construct a whole-genome-based phylogenetic tree with IQ-TREE with 1,000 bootstrap replicates. The best nucleotide substitution model Q.bird+R + F was selected based on the Bayesian Information Criterion (BIC), and was visualized with iTOL [17,26]. FastANI was used to compute the overall Average Nucleotide Identity (ANI), with 95% as the cutoff to define species boundaries [27,28]. DNA-DNA hybridization (DDH) values were calculated using an online tool (ggdc.dsmz.de), and a threshold of 70% was applied for species delineation [29,30].
Genome annotation and pangenome analysis
Genome sequences of reference Bartonella spp. were downloaded from the NCBI RefSeq (https://www.ncbi.nlm.nih.gov/refseq/) [31] (S3 Table). Functional characterization of the Bartonella genomes was conducted using the EggNOG-mapper [32], and virulence factor annotation was conducted using Diamond against the VFDB database [33]. Core genes and unique genes were identified through Roary analysis [34].
Positive selection analysis and genome comparison
After the core gene sequences were obtained, they were translated into coding sequences (CDs) using the Gffread software [35]. The aligned protein sequences, together with their corresponding CDs were processed using PAL2NAL to produce codon-based alignments suitable for selection pressure analysis [36]. The sequence alignments generated by PAL2NAL were used to construct a phylogenetic tree using IQ-TREE [17]. Based on the phylogenetic tree and the aligned sequence, selection pressure analysis was performed using the aBSREL model in HyPhy to identify positively selected genes among the core genes [37]. Positively selected genes were extracted from the core genes using Python scripting. Functional annotation of these positively selected genes was conducted employing EggNOG-mapper [32].
For strains B30 and B17, which were isolated from the same host, genome alignment was performed using Bowtie2 [38]. The alignment results were indexed and single nucleotide polymorphism (SNP) regions were extracted using FreeBayes [39]. The SNP regions were annotated with snpEff [40]. After the identification of SNP sites with significant impact, genes containing these loci were functionally annotated using EggNOG-mapper [32]. The versions and parameter settings of all software used in this study are provided in S5 Table.
Results
Sampling and species identification of bats and their ectoparasites
A total of 28 bats, including Myotis laniger and Myotis adversus, were examined, along with 49 ectoparasites, comprising 34 Penicillidia monoceros, five Nycteribia formosana (processed as ten pooled samples), and ten Spinturnix sp. (processed as 23 pooled samples). Among the bat spleen samples, eight of 28 (28.6%) were positive for Bartonella, including six positives from M. laniger and two from M. adversus. For ectoparasites, nine of 34 P. monoceros (26.5%) were positive, eight of ten pooled N. formosana samples (80.0%) were positive, and one of 23 pooled Spinturnix sp. samples (4.3%) was positive. Percentages indicate the proportion of positive samples among those tested in this study.
Bartonella isolation and molecular identification
A total of 26 strains of Bartonella were isolated from bat spleens and bat ectoparasites, and phylogenetic analysis based on the genes of 16S rRNA, gltA, rpoB, and ftsZ indicated that these isolates represent eight potential novel Bartonella species, which are provisionally designated in this study as Bartonella sp. B10, Bartonella sp. B12, Bartonella sp. B17, Bartonella sp. B23, Bartonella sp. B30, Bartonella sp. B35, Bartonella sp. B39, and Bartonella sp. B41 (Table 1, Fig 1). All gene sequences and whole-genome data have been deposited in GenBank and accession numbers have been obtained (S3 Table).
https://doi.org/10.1371/journal.pntd.0013646.t001
The trees were constructed using the best nucleotide substitution model selected based on the Bayesian Information Criterion (BIC), with K2P+I + G4, TIM3 + F + R3, TN + F + R3, and HKY + F + R2 for 16S rRNA, gltA, rpoB, and ftsZ, respectively. The bootstrap values were 1,000. Strains identified in the present study are shown in bold font, with bootstrap values ≥50 indicated at the nodes. The branch length of the outgroup (Brucella abortus) was reduced by 30% for visualization purposes.
https://doi.org/10.1371/journal.pntd.0013646.g001
Genomic characteristics and phylogenetic analysis
The eight bat-borne Bartonella spp. isolated in this study underwent second and third-generation sequencing. Assembly completeness, estimated from the presence of lineage-specific single-copy marker genes, ranged from 94.0% to 100%, with contamination levels below 5.39%. The genome size of the bat-borne Bartonella spp. ranged from 1.39 to 1.98 Mb, and the GC content ranged from 35.30% to 38.92% (Table 2). Whole genome-based phylogenetic analysis showed that the bat-borne Bartonella spp. formed multiple well-supported monophyletic clades (Fig 2), highlighting their distinct evolutionary lineages. Strains B10, B35, and B41 clustered into a stable branch with full bootstrap support (100%), while strains B17 and B30 also formed a strongly supported monophyletic clade (100% bootstrap support). These two distinct lineages represent ancestral branches to the L4 Bartonella clade. Similarly, strains B23 and B39 consistently grouped together (100% bootstrap support) and were positioned adjacent to a clade containing B. machadoe, B. harrusi, and multiple subspecies of B. vinsonii. In addition, Strain B12 formed a completely distinct branch within the Bartonella lineage L4, without clustering with any of the other strains identified in our study.
https://doi.org/10.1371/journal.pntd.0013646.t002
Strains identified in the present study are shown in bold font, with bootstrap values ≥50 indicated at the nodes.
https://doi.org/10.1371/journal.pntd.0013646.g002
The average nucleotide identity (ANI) values of the bat-borne novel Bartonella spp. were all < 95% (79.76% -91.15%), dDDH values were all < 70% (21.8% -29.5%), confirming that they were distinct Bartonella spp. (S1 Table). The ANI values among the bat-borne Bartonella spp. and their closely related reference Bartonella spp. ranged from 77.55% to 86.21%, all below the 95% threshold, and all dDDH values were all < 70% (Tables 3 and S1), indicating that they were novel Bartonella spp.. https://doi.org/10.1371/journal.pntd.0013646.t003 Genome annotation and pangenome analysis For comparative purposes, four pathogenic Bartonella species with representative pathogenic relevance and high-quality genomes (B. bacilliformis, B. henselae, B. quintana, and B. vinsonii) were included in the analysis. Gene annotation with EggNOG-mapper for the eight bat-borne novel Bartonella spp. of this study and four pathogenic Bartonella spp. showed that the top four most abundant functional categories, in terms of proportion of annotated genes, were J (translation, ribosomal structure and biogenesis), E (amino acid transport and metabolism), L (replication, recombination, and repair), and M (cell wall/membrane/envelope biogenesis) (Figs 3 and S1, S5 Tables). The relative proportion of genes in functional category N (cell motility) in the bat-associated novel Bartonella spp. was significantly lower than that of B. bacilliformis. Moreover, a markedly higher proportion of genes in functional category K (transcription) was observed in strain B39 compared to the other seven Bartonella spp., as well as B. bacilliformis and B. henselae (Figs 3 and S1). The y-axis represents the proportion of annotated genes assigned to each functional category in the Bartonella genomes, while the x-axis denotes different Bartonella species, where B indicates B. bacilliformis KC583, H indicates B. henselae Houston-I, Q indicates B. quintana Toulouse, and V indicates B. vinsonii NCTC12905. https://doi.org/10.1371/journal.pntd.0013646.g003 All eight bat-borne novel Bartonella spp. isolated in this study contained Type IV secretion system (T4SS) and VirB/D4 (Fig 4). Strains B23 and B39, which clustered together in the phylogenetic tree, were observed to exhibit a similar distribution of virulence factors with the pathogenic B. henselae, B. quintana, and B. vinsonii (Fig 4), consistent with the virulence factor profile characteristic of lineage L4, thereby indicating a closely related evolutionary history and pathogenic potential [41]. Solid circles indicate the presence of the virulence factor, while empty circles represent the absence of the virulence factor. https://doi.org/10.1371/journal.pntd.0013646.g004 The core genes, accessory genes, and specific genes of the bat-borne novel Bartonella spp. were identified with pan-genomic analysis (S2 Table). Core functional categories, such as J (Translation, ribosomal structure and biogenesis), C (Energy production and conversion), and K (Transcription), were consistently represented by high gene counts across all species (S2 Table). Positive selection analysis and genome comparison Positive selection branch detection was performed with 1,859 core genes of the eight bat-borne novel Bartonella spp.. When considering only ω values >1, an average of 74 positively selected genes were found in the core genes of the bat-borne novel Bartonella spp., ranging from a minimum of 42 in strain B10 to a maximum of 128 in strain B12. Under the condition of ω value >1 and P value <0.05, fewer genes were identified (e.g., 10 in strain B10 and 2 in strains B12 and B41), whereas no genes met this threshold in strain B23 (S4 Table). While some core genes may experience episodic positive selection, the majority are under strong purifying selection. It should be noted that after stringent multiple testing correction, none of the core genes reached statistically significant levels, highlighting the conservative nature of these essential genes. In this study, the genome sequences of strains B30 and B17, which were isolated from the same host, were compared to identify significant SNP sites. Significant SNP sites in the himD gene were identified through whole-genome comparison (Fig 5B), showing significant genetic differences between the two Bartonella spp.. The HimD is one of the two subunits of the integration host factor [42], a specific DNA-binding protein that plays an important role in gene recombination, transcription, and translation control [43,44]. Compared with 13 reference Bartonella spp., the himD gene in strain B30 was found to have undergone nonsynonymous substitutions at the 25th and 35th amino acid positions (Fig 5A). Compared to strain B17, two nonsynonymous SNPs were identified in the HimD gene of strain B30, resulting in amino acid substitutions at positions 25 (Isoleucine → Valine) and 35 (Threonine → Isoleucine). Furthermore, signals of positive selection were detected in the himD gene of strain B30 (Fig 5B). (A) Amino acid sequence alignments of HimD in strain B30 with 13 other Bartonella HimD. (B) Detailed nucleotide alignments of the himD gene in strains B17 and B30 showing the codon changes responsible for amino acid substitutions at positions 25 and 35. https://doi.org/10.1371/journal.pntd.0013646.g005 Discussion By combining isolation and whole-genome sequencing, the eight novel bat-borne Bartonella spp. were characterized in this study. The sequencing results indicated that the genome length, GC content, and assembly completeness of the new Bartonella species were within the typical ranges observed for the genus (S3 Table). Phylogenetic analyses based on the 16S rRNA, ftsZ, gltA, and rpoB genes consistently supported the placement of the bat-associated Bartonella strains into multiple, well-supported and independent clades, which were in full agreement with the phylogeny reconstructed from concatenated single-copy core genes. Based on the whole-genome-based phylogenetic placement, strain B12 was inferred to represent an independent lineage likely belonging to the broader L4 clade. However, the distribution of virulence factors in strain B12 differed from that of the L4 lineage. Notably, strains B23 and B39 clustered closely with Candidatus Bartonella mayotimonensis in the phylogenetic trees constructed from gltA and rpoB sequences, a species originally identified from a human endocarditis patient and later reported in bats [45,46]. This finding underscores the potential pathogenic relevance of bat-associated Bartonella. In the whole-genome-based phylogeny, major internal nodes received full bootstrap support, with strains B23 and B39 consistently forming a strongly supported pair (100% bootstrap support) adjacent to the L4 clade containing B. machadoe, B. harrusi, and the complex of B. vinsonii subspecies. These results highlight that the newly described bat-associated Bartonella strains expand the phylogenetic diversity of the genus and reveal evolutionary links to clades associated with other mammalian hosts, including rodents, marsupials, and carnivores. Bartonella henselae and B. quintana are two well-known pathogenic Bartonella spp. [3,47]. Among the bat-borne Bartonella species identified in our study, strains B23 and B39 exhibited similar distribution of the virulence factors with B. henselae and B. quintana. The type IV secretion system (T4SS), comprising VirB and associated Bartonella effector proteins (Beps), plays a central role in the pathogenesis of Bartonella species by delivering effector proteins into host cells [41,48]. These effectors modulate host cell signalling, inhibit immune responses, enhance bacterial uptake and intracellular survival, and facilitate other pathogenic processes [41,49]. Both strains B23 and B39 were found to possess complete sets of T4SS-related genes, and together with their phylogenetic positions within the genus Bartonella, this indicates their potential to infect mammalian hosts and exert pathogenic effects. Although previous studies have revealed a high diversity of Bartonella species in bats and bat ectoparasites based on short gene fragments such as gltA, rpoB, or ftsZ [13,50,51], comprehensive whole-genome data of bat-associated Bartonella spp. remain scarce. In this study, a combination of the second and third-generation sequencing methods was used to obtain whole-genome sequences of the novel bat-borne Bartonella spp.. Analysis of COG functional annotations among the bat-associated Bartonella species core genes revealed a highly conserved distribution pattern across most categories. Compared to other strains, a higher proportion of genes in functional category K (transcription) was shown by 311-B39, suggesting a potentially richer and more active gene expression during transcription. [48,52]. Additionally, several important SNP sites were identified, including a mutation site in the himD gene, and positive selection analysis revealed the himD gene as an episodically positively selected gene in strain B30. The HimD gene encodes the integration host factor (IHF) subunit, and it was demonstrated that IHF assisted in regulating chromosome function in bacteria by influencing the morphology and organisation of DNA [53]. The transcription of type III secretion system (T3SS) genes and the replication of the pYGK plasmid were also shown to require the involvement of IHF [54,55]. Moreover, the integration host factor has been demonstrated to be essential for controlling virulence gene expression in several bacterial species, including Salmonella, Shigella, and Brucella [56–59]. Positive selection of the IHFβ subunit in strain B30 may be related to its adaptation to the host environment. Under immune or other environmental pressures, changes in the IHFβ subunit may aid strain B30 in evading host defense mechanisms, enhancing its survival and colonization within the host [60]. Strain B30, which was detected exclusively in Penicillidia monoceros samples, may reflect adaptive pressures associated with its specialized evolutionary adaptation to bat ectoparasites. In contrast, strain B17 was detected in both ectoparasites and two bat species, whereas strains B39 and B41 were detected only in bat species. Future researches are needed to explore the impact of genetic variation on the function of IHFβ, as well as its effect on host specificity. In conclusion, with the availability of the whole-genome sequence of bat-borne Bartonella spp., our study significantly enhanced the understanding of Bartonella spp. diversity and offered novel perspectives on the pathogenicity, virulence mechanisms, and evolutionary adaptations of bat-borne Bartonella spp.. Conclusions Through isolation and whole-genome sequencing, eight novel bat-associated Bartonella species were identified. Their distinct phylogenetic positions, virulence gene profiles, and functional genomic features revealed previously unrecognized diversity within the genus. Strains B23 and B39 were found to carry a complete set of T4SS genes, indicating potential pathogenicity. In strain B30, a mutation site in the himD gene suggested possible adaptation to a specific host environment. Our findings provided important genomic insights into the evolution, host specificity, and virulence potential of bat-associated Bartonella species, highlighting the necessity for continued surveillance and functional investigations. Supporting information S1 Fig. Supplementary figure of the COG database annotation results. https://doi.org/10.1371/journal.pntd.0013646.s001 (TIF) S1 Table. ANI values and dDDH values of the eight bat-borne novel Bartonella spp. of this study. https://doi.org/10.1371/journal.pntd.0013646.s002 (PDF) S2 Table. Statistical results of pan-genome analysis and core gene functional annotation of the eight bat-borne novel Bartonella spp. https://doi.org/10.1371/journal.pntd.0013646.s003 (PDF) S3 Table. Genomic and gene sequence information (16S rRNA, gltA, ftsZ, rpoB) with accession numbers. https://doi.org/10.1371/journal.pntd.0013646.s004 (PDF) S4 Table. Episodic positive selection analysis of core genes in bat-borne Bartonella was conducted. https://doi.org/10.1371/journal.pntd.0013646.s005 (PDF) S5 Table. Functional categories of COG protein function annotations and the software versions with parameter settings used in this study. https://doi.org/10.1371/journal.pntd.0013646.s006 (PDF) S6 Table. Primers used for PCR in this study. https://doi.org/10.1371/journal.pntd.0013646.s007 (PDF) S1 File. Raw PCR positivity data and associated metadata for Bartonella detection in bats and their ectoparasites. This dataset provides sample-level PCR positivity data for bats and their ectoparasites. The table includes the following variables: Sample ID, Animal ID, Latitude, Longitude, Sample collection method, Sample material, Host identification, Detection target, Detection method, Detection outcome, and Species identification. PCR results are represented as “1” for positive and “0” for negative detections. Missing values are indicated by blank cells. https://doi.org/10.1371/journal.pntd.0013646.s008 (XLSX) References 1. Chomel BB, Boulouis H-J, Breitschwerdt EB, Kasten RW, Vayssier-Taussat M, Birtles RJ, et al. Ecological fitness and strategies of adaptation of Bartonella species to their hosts and vectors. Vet Res. 2009;40(2):29. pmid:19284965 2. Brenner DJ, O’Connor SP, Winkler HH, Steigerwalt AG. Proposals to unify the genera Bartonella and Rochalimaea, with descriptions of Bartonella quintana comb. nov., Bartonella vinsonii comb. nov., Bartonella henselae comb. nov., and Bartonella elizabethae comb. nov., and to remove the family Bartonellaceae from the order Rickettsiales. Int J Syst Bacteriol. 1993;43(4):777–86. pmid:8240958 3. Clarridge JE 3rd, Raich TJ, Pirwani D, Simon B, Tsai L, Rodriguez-Barradas MC, et al. Strategy to detect and identify Bartonella species in routine clinical laboratory yields Bartonella henselae from human immunodeficiency virus-positive patient and unique Bartonella strain from his cat. J Clin Microbiol. 1995;33(8):2107–13. pmid:7559957 4. Daly JS, Worthington MG, Brenner DJ, Moss CW, Hollis DG, Weyant RS, et al. Rochalimaea elizabethae sp. nov. isolated from a patient with endocarditis. J Clin Microbiol. 1993;31(4):872–81. pmid:7681847 5. Kerkhoff FT, Bergmans AM, van Der Zee A, Rothova A. Demonstration of Bartonella grahamii DNA in ocular fluids of a patient with neuroretinitis. J Clin Microbiol. 1999;37(12):4034–8. pmid:10565926 6. Dehio C. Bartonella-host-cell interactions and vascular tumour formation. Nat Rev Microbiol. 2005;3(8):621–31. pmid:16064054 7. Gutiérrez R, Shalit T, Markus B, Yuan C, Nachum-Biala Y, Elad D, et al. Bartonella kosoyi sp. nov. and Bartonella krasnovii sp. nov., two novel species closely related to the zoonotic Bartonella elizabethae, isolated from black rats and wild desert rodent-fleas. Int J Syst Evol Microbiol. 2020;70(3):1656–65. pmid:32100689 8. Sato S, Kabeya H, Fujinaga Y, Inoue K, Une Y, Yoshikawa Y, et al. Bartonella jaculi sp. nov., Bartonella callosciuri sp. nov., Bartonella pachyuromydis sp. nov. and Bartonella acomydis sp. nov., isolated from wild Rodentia. Int J Syst Evol Microbiol. 2013;63(Pt 5):1734–40. pmid:22941296 9. Han H-J, Wen H, Zhou C-M, Chen F-F, Luo L-M, Liu J, et al. Bats as reservoirs of severe emerging infectious diseases. Virus Res. 2015;205:1–6. pmid:25997928 10. McKee CD, Bai Y, Webb CT, Kosoy MY. Bats are key hosts in the radiation of mammal-associated Bartonella bacteria. Infect Genet Evol. 2021;89:104719. pmid:33444855 11. Szentiványi T, Takács N, Sándor AD, Péter Á, Boldogh SA, Kováts D, et al. Bat-associated ticks as a potential link for vector-borne pathogen transmission between bats and other animals. PLoS Negl Trop Dis. 2024;18(10):e0012584. pmid:39453968 12. Han H-J, Wen H-L, Zhao L, Liu J-W, Luo L-M, Zhou C-M, et al. Novel Bartonella species in insectivorous bats, Northern China. PLoS One. 2017;12(1):e0167915. pmid:28081122 13. Han H-J, Li Z-M, Li X, Liu J-X, Peng Q-M, Wang R, et al. Bats and their ectoparasites (Nycteribiidae and Spinturnicidae) carry diverse novel Bartonella genotypes, China. Transbound Emerg Dis. 2022;69(4):e845–58. pmid:34695291 14. Deng HK, Le Rhun D, Lecuelle B, Le Naour E, Vayssier-Taussat M. Role of the spleen in Bartonella spp. infection. FEMS Immunol Med Microbiol. 2012;64(1):143–5. pmid:22098417 15. Rao H, Li S, Lu L, Wang R, Song X, Sun K, et al. Genetic diversity of Bartonella species in small mammals in the Qaidam Basin, western China. Sci Rep. 2021;11(1):1735. pmid:33462399 16. Evans NJ, Brown JM, Demirkan I, Singh P, Getty B, Timofte D, et al. Association of unique, isolated treponemes with bovine digital dermatitis lesions. J Clin Microbiol. 2009;47(3):689–96. pmid:19144804 17. Nguyen L-T, Schmidt HA, von Haeseler A, Minh BQ. IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol Biol Evol. 2015;32(1):268–74. pmid:25371430 18. Chen S. Ultrafast one-pass FASTQ data preprocessing, quality control, and deduplication using fastp. Imeta. 2023;2(2):e107. pmid:38868435 19. Chen S, Zhou Y, Chen Y, Gu J. fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics. 2018;34(17):i884–90. pmid:30423086 20. Prjibelski A, Antipov D, Meleshko D, Lapidus A, Korobeynikov A. Using SPAdes de novo assembler. Curr Protoc Bioinformatics. 2020;70(1):e102. pmid:32559359 21. Simão FA, Waterhouse RM, Ioannidis P, Kriventseva EV, Zdobnov EM. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics. 2015;31(19):3210–2. pmid:26059717 22. Parks DH, Imelfort M, Skennerton CT, Hugenholtz P, Tyson GW. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 2015;25(7):1043–55. pmid:25977477 23. Seemann T. Prokka: rapid prokaryotic genome annotation. Bioinformatics. 2014;30(14):2068–9. pmid:24642063 24. Rozewicki J, Li S, Amada KM, Standley DM, Katoh K. MAFFT-DASH: integrated protein sequence and structural alignment. Nucleic Acids Res. 2019;47(W1):W5–10. pmid:31062021 25. Emms DM, Kelly S. OrthoFinder: phylogenetic orthology inference for comparative genomics. Genome Biol. 2019;20(1):238. pmid:31727128 26. Letunic I, Bork P. Interactive Tree of Life (iTOL) v6: recent updates to the phylogenetic tree display and annotation tool. Nucleic Acids Res. 2024;52(W1):W78–82. pmid:38613393 27. Goris J, Konstantinidis KT, Klappenbach JA, Coenye T, Vandamme P, Tiedje JM. DNA-DNA hybridization values and their relationship to whole-genome sequence similarities. Int J Syst Evol Microbiol. 2007;57(Pt 1):81–91. pmid:17220447 28. Ciufo S, Kannan S, Sharma S, Badretdin A, Clark K, Turner S, et al. Using average nucleotide identity to improve taxonomic assignments in prokaryotic genomes at the NCBI. Int J Syst Evol Microbiol. 2018;68(7):2386–92. pmid:29792589 29. Meier-Kolthoff JP, Carbasse JS, Peinado-Olarte RL, Göker M. TYGS and LPSN: a database tandem for fast and reliable genome-based classification and nomenclature of prokaryotes. Nucleic Acids Res. 2022;50(D1):D801–d7. pmid:34634793 30. Meier-Kolthoff JP, Auch AF, Klenk H-P, Göker M. Genome sequence-based species delimitation with confidence intervals and improved distance functions. BMC Bioinform. 2013;14:60. pmid:23432962 31. O’Leary NA, Wright MW, Brister JR, Ciufo S, Haddad D, McVeigh R, et al. Reference sequence (RefSeq) database at NCBI: current status, taxonomic expansion, and functional annotation. Nucleic Acids Res. 2016;44(D1):D733-45. pmid:26553804 32. Huerta-Cepas J, Szklarczyk D, Heller D, Hernández-Plaza A, Forslund SK, Cook H, et al. eggNOG 5.0: a hierarchical, functionally and phylogenetically annotated orthology resource based on 5090 organisms and 2502 viruses. Nucleic Acids Res. 2019;47(D1):D309–14. pmid:30418610 33. Buchfink B, Reuter K, Drost H-G. Sensitive protein alignments at tree-of-life scale using diamond. Nat Methods. 2021;18(4):366–8. pmid:33828273 34. Page AJ, Cummins CA, Hunt M, Wong VK, Reuter S, Holden MTG, et al. Roary: rapid large-scale prokaryote pan genome analysis. Bioinformatics. 2015;31(22):3691–3. pmid:26198102 35. Pertea G, Pertea M. GFF utilities: GffRead and GffCompare. F1000Res. 2020;9. pmid:32489650 36. Suyama M, Torrents D, Bork P. PAL2NAL: robust conversion of protein sequence alignments into the corresponding codon alignments. Nucleic Acids Res. 2006;34(Web Server issue):W609–12. pmid:16845082 37. Pond SLK, Frost SDW, Muse SV. HyPhy: hypothesis testing using phylogenies. Bioinformatics. 2005;21(5):676–9. pmid:15509596 38. Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012;9(4):357–9. pmid:22388286 39. Garrison Erik GM. Haplotype-based variant detection from short-read sequencing. 2012. 40. Cingolani P, Platts A, Wang LL, Coon M, Nguyen T, Wang L, et al. A program for annotating and predicting the effects of single nucleotide polymorphisms, SnpEff: SNPs in the genome of Drosophila melanogaster strain w1118; iso-2; iso-3. Fly (Austin). 2012;6(2):80–92. pmid:22728672 41. Wagner A, Dehio C. Role of distinct type-IV-secretion systems and secreted effector sets in host adaptation by pathogenic Bartonella species. Cell Microbiol. 2019;21(3):e13004. pmid:30644157 42. Flamm EL, Weisberg RA. Primary structure of the hip gene of Escherichia coli and of its product, the beta subunit of integration host factor. J Mol Biol. 1985;183(2):117–28. pmid:3159903 43. Muir RE, Gober JW. Role of integration host factor in the transcriptional activation of flagellar gene expression in Caulobacter crescentus. J Bacteriol. 2005;187(3):949–60. pmid:15659673 44. Kasho K, Katayama T. DnaA binding locus datA promotes DnaA-ATP hydrolysis to enable cell cycle-coordinated replication initiation. Proc Natl Acad Sci U S A. 2013;110(3):936–41. pmid:23277577 45. Lilley TM, Wilson CA, Bernard RF, Willcox EV, Vesterinen EJ, Webber QMR, et al. Molecular detection of candidatus Bartonella mayotimonensis in North American bats. Vector Borne Zoonotic Dis. 2017;17(4):243–6. pmid:28165925 46. Veikkolainen V, Vesterinen EJ, Lilley TM, Pulliainen AT. Bats as reservoir hosts of human bacterial pathogen, Bartonella mayotimonensis. Emerg Infect Dis. 2014;20(6):960–7. pmid:24856523 47. Okaro U, Addisu A, Casanas B, Anderson B. Bartonella species, an emerging cause of blood-culture-negative endocarditis. Clin Microbiol Rev. 2017;30(3):709–46. pmid:28490579 48. Engel P, Salzburger W, Liesch M, Chang C-C, Maruyama S, Lanz C, et al. Parallel evolution of a type IV secretion system in radiating lineages of the host-restricted bacterial pathogen Bartonella. PLoS Genet. 2011;7(2):e1001296. pmid:21347280 49. Vayssier-Taussat M, Le Rhun D, Deng HK, Biville F, Cescau S, Danchin A, et al. The Trw type IV secretion system of Bartonella mediates host-specific adhesion to erythrocytes. PLoS Pathog. 2010;6(6):e1000946. pmid:20548954 50. McKee CD, Kosoy MY, Bai Y, Osikowicz LM, Franka R, Gilbert AT, et al. Diversity and phylogenetic relationships among Bartonella strains from Thai bats. PLoS One. 2017;12(7):e0181696. pmid:28727827 51. Nabeshima K, Sato S, Brinkerhoff RJ, Amano M, Kabeya H, Itou T, et al. Prevalence and genetic diversity of Bartonella Spp. in Northern Bats (Eptesicus nilssonii) and their blood-sucking Ectoparasites in Hokkaido, Japan. Microb Ecol. 2023;85(1):298–306. pmid:34981145 52. Saenz HL, Engel P, Stoeckli MC, Lanz C, Raddatz G, Vayssier-Taussat M, et al. Genomic analysis of Bartonella identifies type IV secretion systems as host adaptability factors. Nat Genet. 2007;39(12):1469–76. pmid:18037886 53. Sarkar T, Petrov AS, Vitko JR, Santai CT, Harvey SC, Mukerji I, et al. Integration host factor (IHF) dictates the structure of polyamine-DNA condensates: implications for the role of IHF in the compaction of bacterial chromatin. Biochemistry. 2009;48(4):667–75. pmid:19132923 54. Childress C, Feuerbacher LA, Phillips L, Burgum A, Kolodrubetz D. Mlc is a transcriptional activator with a key role in integrating cyclic AMP receptor protein and integration host factor regulation of leukotoxin RNA synthesis in Aggregatibacter actinomycetemcomitans. J Bacteriol. 2013;195(10):2284–97. pmid:23475968 55. Waite C, Schumacher J, Jovanovic M, Bennett M, Buck M. Negative autogenous control of the master type iii secretion system regulator HrpL in Pseudomonas syringae. mBio. 2017;8(1):e02273-16. pmid:28119474 56. Porter ME, Dorman CJ. Positive regulation of Shigella flexneri virulence genes by integration host factor. J Bacteriol. 1997;179(21):6537–50. pmid:9352898 57. Mangan MW, Lucchini S, Danino V, Cróinín TO, Hinton JCD, Dorman CJ. The integration host factor (IHF) integrates stationary-phase and virulence gene expression in Salmonella enterica serovar Typhimurium. Mol Microbiol. 2006;59(6):1831–47. pmid:16553887 58. Pozdeev G, Beckett MC, Mogre A, Thomson NR, Dorman CJ. Reciprocally rewiring and repositioning the Integration Host Factor (IHF) subunit genes in Salmonella enterica serovar Typhimurium: impacts on physiology and virulence. Microb Genom. 2022;8(2):000768. pmid:35166652 59. Sieira R, Comerci DJ, Pietrasanta LI, Ugalde RA. Integration host factor is involved in transcriptional regulation of the Brucella abortus virB operon. Mol Microbiol. 2004;54(3):808–22. pmid:15491369 60. Schmidt K, Santos-Matos G, Leopold-Messer S, El Chazli Y, Emery O, Steiner T, et al. Integration host factor regulates colonization factors in the bee gut symbiont Frischella perrara. Elife. 2023;12:e76182. pmid:37057993

webintern@dakdan.com