Skip to main content

Comprehensive molecular characterization of complete mitogenome assemblies of 33 Eimeria isolates infecting domestic chickens

Abstract

Background

Coccidiosis caused by Eimeria is one of the most severe chicken diseases and poses a great economic threat to the poultry industry. Understanding the evolutionary biology of chicken Eimeria parasites underpins development of new interactions toward the improved prevention and control of this poultry disease.

Methods

We presented an evolutionary blueprint of chicken coccidia by genetically characterizing complete mitogenome assemblies of 33 isolates representing all seven known Eimeria species infecting chickens in China. Further genome- and gene-level phylogenies were also achieved to better understand the evolutionary relationships of these chicken Eimeria at the species level.

Results

33 mitogenomes of chicken eimerian parasites ranged from 6148 bp to 6480 bp in size and encoded typical mitochondrial compositions of apicomplexan parasites including three protein-coding genes (PCGs), seven fragmented small subunit (SSU) and 12/13 fragmented large subunit (LSU) rRNAs. Comparative genomics provided an evolutionary scenario for the genetic diversity of PCGs-cytochrome c oxidase subunits 1 and 3 (cox1 and cox3) and cytochrome b (cytb); all were under purifying selection with cox1 and cox3 being the lowest and highest evolutionary rates, respectively. Genome-wide phylogenies classified the 33 Eimeria isolates into seven subgroups, and furthermore Eimeria tenella and Eimeria necatrix were determined to be more closely related to each other than to the other eight congenic species. Single/concatenated mitochondrial protein gene-based phylogenies supported cox1 as the genetic marker for evolutionary and phylogenetic studies for avain coccidia.

Conclusions

To our knowledge, these are the first population-level mitogenomic data on the genus Eimeria, and its comprehensive molecular characterization provides valuable resources for systematic, population genetic and evolutionary biological studies of apicomplexan parasites in poultry.

Graphical Abstract

Background

Chicken coccidiosis, caused by apicomplexan parasites of the genus Eimeria, is a widespread and highly pathogenic avian disease posing a significant threat to the poultry industry worldwide [1, 2]. There are seven species of Eimeria responsible for morbidity and socioeconomic burdens including Eimeria necatrix, E. tenella, E. maxima, E. praecox, E. mitis, E. brunetti and E. acervulina [3]. Besides, three cryptic Eimeria operational taxonomic units (OTUs) were recently identified as endemic to Australian chicken populations and later given the names Eimeria lata, E. nagambie and E. zaria based on their genotypic and phenotypic properties [4,5,6]. Infections with Eimeria spp. can damage the digestive tract of chickens and lead to malabsorption of nutrients and lethal diarrhea, which is reflected in clinical symptoms: poor weight gain, low egg production and even death [1, 3]. There are many reports of coccidiosis prevalence in countries worldwide, with 70–92% prevalence in Romania (92%), Greece (86%), Brazil (91%), India (81%), Algeria (72%), Turkey (70%) and China (88%), and multi-Eimeria species infections predominate [7,8,9,10,11,12,13,14]. The annual economic loss for chicken farmers has been estimated at over £10.4 billion worldwide [15].

Because of similar oocyst morphotypes and overlapping biological features of chicken Eimeria spp. [16, 17], species identification based on morphological characteristics is difficult. Therefore, there is an urgent need to develop an efficient approach to identify Eimeria infections for clinical diagnosis; moreover, phylogenetic relationships and biogeographic range evolution of diverse strains can also provide valuable information to support diagnosis and control of chicken coccidiosis. Therefore, it might be possible to achieve this goal by utilizing next-generation sequencing (NGS) approach-based species identification, phylogenetic constructions and population genetic analyses. Encouragingly, Vermeulen and colleagues have established a ribosomal protein 18 (18S)-based NGS methodology to assess Eimeria communities of the Australian marsupial [18]. Hinsu et al. have also reported the application of Illumina MiSeq deep sequencing to 18S amplicons of chicken Eimeria and achieved detection of all seven validated species including their cryptic genotypes [19]. In addition, increased genetic evidence showed that the taxonomy and diversity of certain taxa or specific groups of chicken Eimeria can be equally revealed by using the mitochondrial (mt) gene-based NGS owing to matrilineal inheritance, absent recombination and rapid evolution rate of these molecules [20,21,22]. For example, Hauck et al. showed that using NGS to sequence mt cytochrome c oxidase subunit 1 (cox1) was efficient for identification of Eimeria species that circulated in chicken flocks [13]. Similarly, Snyder et al. also established the mt cytochrome c oxidase subunit 3 (cox3) amplicon library-based NGS pipeline analysis to differentiate E. acervulina, E. tenella, E. maxima, E. mitis, E. necatrix and E. praecox from mixed infected samples [23]. Nevertheless, it appears noteworthy that the single gene/locus-based NGS approach fails to provide sufficient genetic information to understand variations at the species and genus levels. Consequently, using NGS-based mitogenomic data becomes an ideal candidate strategy because it can not only provide comprehensive molecular insights into intra-/interspecific variability but also yields a broad taxonomy-range phylogenetic and evolutionary justification for Eimeria parasites [24, 25].

Herein, we sequenced complete mitogenomes of 33 Eimeria isolates infecting domestic chickens in China and genetically characterized these isolates by comparisons with other chicken eimerian parasites for which mitogenomic datasets are available in public databases. Furthermore, genome- and gene-level phylogenetic analyses were achieved to better understand the evolutionary relationships of chicken Eimeria at the species level. This comprehensive molecular information will provide valuable resources for systematic, population genetic and evolutionary biological studies of Eimeria parasites in poultry.

Methods

Litter sampling, parasite analysis and DNA extraction

From 2019 to 2021, 153 commercial farms that were in the finishing phase of the productive cycle were selected and sampled as part of ongoing surveys of coccidia on commercial broiler farms in China: 32 in Sichuan, 30 in Henan, 25 in Anhui, 23 in Chongqing, 18 in Guangdong, 15 in Fujian and 10 in Zhejiang. Litter samples were randomly collected from each shed while walking in a “zigzag” pattern [26]. About 250-g litter from each sampling was bagged and transported to the laboratory where it was kept at 4 °C. Oocysts per gram (OPG) of each litter was quantitated using a McMaster chamber. For species identification, positive litter samples were mashed and suspended with running water, followed by filtration through a series of sieves (212, 180, 75 and 45 mm; Endecotts, London, UK) and culture in a 5% (w/v) potassium dichromate solution for 72 h at 28 °C under forced aeration. Oocysts were sporulated and microscopically speciated using morphometric keys [27, 28]. Furthermore, pure sporulated oocysts from 33 isolates that represent all known Eimeria species infecting chickens were enriched with saturated sodium nitrate flotation technique. One hundred sporulated oocysts from each isolate were used for genomic DNA extraction. After vortex breakdown, the total genomic DNA was extracted from oocysts using the Genomic DNA Kit (TIANGEN, Beijing, China). Following the PCR amplification using Eimeria-specific primers ERIB1 (forward: 5′-ACCTGGTTGATCCTGCCAG-3′) and ERIB10 (reverse: 5′-CTTCCGCAGGTTCACCTACGG-3′) [29] and comparing with the targeted 18S regions, those isolates were determined to share > 99.5% identities with query sequences in GenBank nos. KT184333 (E. acervulina), KT184337 (E. brunetti), KT184346 (E. maxima), KT184349 (E. necatrix), KT184351 (E. praecox), KT184354 (E. tenella) and FR775307 (E. mitis).

Genome sequencing, assembly and annotation

After species identification, approximately 2 million sporulated oocysts from each of the 33 isolates were used for mitogenome sequencing. Following quality and quantity assessment, ~ 3 µg high-quality DNA from each sample was fragmented to construct 350-bp paired-end (PE) libraries and sequenced on an Illumina HiSeq X-TEN platform (BerryGenomics, Beijing, China). The clean reads (~ 1.8 Gb) were assembled with MITObim [30] using the available mitogenomes of other Eimeria species from GenBank (GenBank nos. JN864949, KX094949, KX094954, HQ702479, HQ702480, HQ702481, HQ702483 and KC409031) as the references. These assembled mitogenomes were also validated by PCR amplifications using four overlapping fragments. These four overlapping fragments were located between cytochrome b (cytb) and cox1 (~ 2.0 kb), between cox1 and large subunit A (LSUA) (~ 3.0 kb), between LSUA and large subunit E (LSUE) (~ 2.0 kb) and between LSUE and cytb (~ 1.0 kb), respectively (Fig. 1), and their corresponding PCR primers were designed based on the aforementioned mitogenome sequences of Eimeria species and are shown in Additional file 1: Table S1. PCR reactions were achieved using a 50-μl reaction volume containing 3 μl genomic DNA (≥ 10 ng), 25 μl 2 × HiFi TransTaq PCR SuperMix (TaKaRa, Biotech, Dalian, China), 1 µl sense primer (10 pmol), 1 µl anti-sense primer (10 pmol) and 20 µl nuclease-free water. The PCR conditions were: 95 °C for 5 min denaturation, 95 °C for 45 s for 30 cycles and 45 ~ 55 °C for 45 s, and 68 °C for 1 ~ 3 min according to the Tm values and the product lengths, and a final extension at 68 °C for 10 min. After agarose gel detections, all target amplicons were column-purified and sequenced either directly or following sub-cloning into the pMD19-T vector (TaKaRa, Biotech, Dalian, China). Each amplicon was sequenced three times to ensure maximum accuracy. Mitogenome annotation was performed with MITOS [31] and manual inspection based on a whole genome-guided alignment using poultry Eimeria spp. for which complete mitogenome sequences were available so far.

Fig. 1
figure 1

Structure views of mitogenomes of 33 chicken Eimeria isolates identified in this study. a–g Mitochondrial genome organizations of the Eimeria isolates 1–33 corresponding to seven chicken-infecting Eimeria species. Isolates in the same species are arranged together and contrast with the reported mitogenome organization. Red color indicates protein-coding genes, blue color indicates fragments of rRNA genes, green and purple colors indicate positive and negative GC-skew, respectively, and black color indicates the GC content of every position in the genome. h The average AT-skew and GC-skew values of the complete mitogenomes (Entire), concatenated PCG datasets (PCGs), LSU and SSU rRNAs of each species

Sequence characterization

Nucleotide composition and codon usage of 33 eimerian mitogenomes were measured using Geneious and PhyloSuite [32, 33]. The nucleotide skewness of mitogenomes was performed using the formulas [34]: AT skew = (A − T)/(A + T) and GC skew = (G − C)/(G + C). Nucleotide and amino acid divergences of protein-coding genes (PCGs) were determined using DNAstar [35]. The multi-alignment of PCGs obtained from 33 sequenced eimerian isolates and available chicken eimerian parasites including newly recognized E. lata, E. nagambie and E. zaria were separately aligned using MEGA software [36]. Based on those alignments, the sliding window analysis was used to compute the nucleotide diversity Pi (π) using a 200-bp window and 20-bp steps, followed by assessment of genetic structure using the Wright’s fixation index (FST) with 1000 replicates as a permutation test in DnaSP ver. 5.10 [37]. In addition, the evolutionary rate and the ratio of the nonsynonymous substitution (Ka) and synonymous substitution (Ks) of each PCG were calculated using KaKs_Calculator [38]. Genetic distances among these 10 Eimeria species mitogenomes were calculated based on Kimura-2-parameter (K2P) [39] with MEGA X.

Phylogenetic characterization

The phylogenies were reconstructed based on the mitogenomic datasets of 33 Eimeria isolates and other related species (Additional file 2: Table S2). Nucleotide sequence alignments were yielded from the complete mitogenomes, concatenated datasets of dual/triple PCGs or individual PCG of coccidian parasites using MAFFT ver. 7.271 [40], followed by filtration of the ambiguous regions with GBLOCKS ver. 0.91b [41]. Three algorithms including the maximum parsimony (MP), maximum-likelihood (ML) and Bayesian inference (BI) were used to reconstruct the phylogenetic relationships of chicken Eimeria species, and Isospora sp. (GenBank no. KP658103) was treated as the outgroup and included in each analysis. For the MP analysis, the complete nucleotide sequence dataset of mitogenomes or multi-alignments of the nucleotide sequence for the single, dual or triple PCGs were analyzed using equally weighted parsimony and heuristic searches with a tree-bisection-reconnection (TBR) branch-swapping in PAUP* [42]. One thousand replicates of Wagner trees (using random addition sequences) were chosen, and five trees of each replication were saved, followed by obtaining the optimal topology with the Kishino-Hasegawa method. Bootstrap resampling with 1000 replications was computed for each nodal support. The ML analysis was implemented with PHYML ver. 3.0.1 [43] using the optimal evolutionary models “TIM + F + I + G4” for the complete mitogenomic dataset and “GTR + F + I + G4” for the dual and triple concatenated PCGs and individual PCG datasets that were selected with the “Auto” option on W-IQ-TREE web server (http://iqtree.cibiv.univie.ac.at) according to Bayesian information criterion (BIC). ML trees were reconstructed with a 10,000-replicate and an ultra-fast bootstrap approximation. Within the BI trees, the optimal evolutionary model “CAT + GTR + G” was selected for all mtDNA datasets using ModelFinder [44]. The BI analyses were carried out using MrBayes ver. 3.2.7 [45], with four independent Markov chains, running for 1,000,000 (complete mitogenomic dataset) and 100,000 (dual and triple concatenated PCGs and individual PCG datasets) metropolises coupled with Monte Carlo generations, sampling a tree every 0.1% generations. The first 25% of the trees were eliminated as “burn-in” when the average standard deviation (SD) of the split frequencies decreased to < 0.01, and the remaining trees were used to calculate Bayesian posterior probabilities (PPs). The evolutionary distance was estimated using the MrBayes order (aamodelpr = mixed) with default parameters. A consensus tree was obtained and visualized using TreeviewX (https://www.linuxlinks.com/treeviewx/).

Results and discussion

General features of 33 Eimeria mitogenomes

The complete mitogenomes of 33 eimerian isolates representing all seven species known to infect chickens in China varied from 6148 to 6480 bp in size (GenBank accession nos. OP800493–OP800525; Table 1). Each genome encoded three mt PCGs including cox1, cox3 and cytb as well as seven fragmented small subunit (SSU) and 12 fragmented LSU rRNAs, except for E. mitis isolates, which contained another 56-bp LSU rRNA (LSU15) located between LSU13 and LSUD (Fig. 1a–g). The overall base composition of each mitogenome consistently exhibited high AT bias (64.52–67.37%) with T being the most favored base and G the least favored, similar to those observed in other Eimeria members [46,47,48,49,50].

Table 1 List of Chinese Eimeria isolates sequenced in the present study

Within SSU rRNAs, seven gene fragments including SSUA, SSUF, SSUD, SSU9, SSU8, SSUB and SSUE ranged from 37 (SSUF) to 116 bp (SSUB) in size. For LSU rRNAs, 12 gene fragments including LSUF, LUSG, LUSC, LSU10, LSU13, LSUD, LSU2, LSUA, LSU1, LUSB, LSU3 and LSUE ranged from 16 (LUSC) to 188 bp (LSUE) in size, very similar to those of other Eimeria species reported to date [25, 46,47,48,49,50]. Furthermore, AT-skew and GC-skew values of rRNAs are shown in Fig. 1h, and it was clear that these mitogenomes shared a same skewness pattern: a slightly higher A than T (except for LSU rRNAs) in E. necatrix (− 0.003), E. maxima (− 0.021) and E. brunetti (− 0.008) and a slightly lower G than C (except for SSU rRNAs) in E. necatrix (− 0.021) and E. tenella (− 0.029).

For three PCGs, cox1, cox3 and cytb located between cytb and LSUF, LSUA and LSU1 and SSUA and cox1, respectively, with an obvious bias towards AT (Fig. 1). Such AT bias was also reflected in their RSCU and codon usage patterns of PCGs (Fig. 2); for instance, within the initiation codon choice, the cox1 gene was inferred to start with codons TTG, ATN (ATT and ATG) and GTN (GTG and GTT), the cox3 gene was inferred to start with codons ATT and TTA, and the cytb gene was inferred to start with codon ATG (Fig. 2h). Correspondingly, the standard stop codon TAG was used to terminate the cox1 and cox3 genes and the cytb gene used TAA as the stop codon, consistent with the previous study [49]. Moreover, RSCU comparisons showed that these 33 eimerian isolates consistently used TAA as the stop codon, except for E. acervulina isolates, which used TGA and TAA as stop codons (Fig. 2), inconsistent with mitogenomes of some other Apicomplexa parasites in which an abbreviated stop codon (TA or T) was used as the stop codon [50]. It was also evident that the most frequently used codon of these 33 eimerian isolates was AGA (RSCU ranging from 4.08–4.86), followed by UUA (2.67–3.65), GGU (2.44–2.62) and CCA (1.95–2.67). These codon patterns were similarly reflected in the amino acid usage frequency of PCGs, and the most frequently used amino acid was Leu (count ranging from 166–171), followed by Ser (102–112), Phe (98–104) and Ile (93–108) (Fig. 2a-g).

Fig. 2
figure 2

Codon usage in PCGs of mitogenomes of 33 chicken Eimeria isolates studied here at the species level. a–g RSCU and the number of codons of PCG genes used in each species based on their corresponding isolates. Codon families are plotted under the X axis and represented by different color bars, and the codon numbers are indicated by the black line graph (right axis scale). h Start and stop codon usages of three PCG genes in all isolate mitogenomes

Mitogenome diversity

Most gene sizes of mitogenomes of these 33 Eimeria isolates were consistent with those of other congeneric species (Additional file 3: Table S3). Nevertheless, to further understand the evolutionary divergence among the genus Eimeria, the intra- and interspecific variability was determined across the 33 isolates using the nucleotide and amino acid sequences. The interspecific divergences were determined to range from 1.5 to 12.8% while intraspecific differences were < 0.3% (Additional file 4: Table S4). Both ranges were lower than those estimated in Australian chicken Eimeria by Morgan and Godwin [25], to some extent supporting a more general systematic congruence among Chinese chicken eimerian parasites sequenced in this study. Meanwhile, it seemed that the lowest interspecific divergence always occurred between E. tenella and E. necatrix, suggesting their closer genetic similarity among Eimeria species. Furthermore, divergences were also focused on various portions of the mitogenomes. For rRNAs, the nucleotide sequence divergences of SSU and LSU rRNAs ranged from 0.2 to 4.6% and 0.4 to 8.2%, respectively; within PCG genes, the nucleotide sequence divergences ranged from 2.1 to 17.6%, and amino acid sequence divergences ranged from 0.8 to 10.1%, with cox3 being the most variable gene, in line with previous findings in species of Plasmodium, Theileria and Babesia [51, 52].

To confirm the aforementioned sequence diversities within and between mt genes, sliding window analysis was also carried out by a combination of the 33 Eimeria isolates and other congeneric species, with a comparison between Chinese and Australian Eimeria populations. As shown in Fig. 3a–g, the intraspecific Pi values of seven Eimeria species, regardless of sources (China or Australia), all were relatively low (0 to 0.069) with a significantly increased trend in the following order: E. acervulina < E. tenella < E. necatrix < E. praecox < E. maxima < E. brunetti < E. mitis. It was significant that various numbers of peaks presented within comparisons of interspecific diversity depending on species pairs, suggesting that there is still a considerable number of alternative genetic loci to be determined as species-specific markers for identification and differentiation of chicken coccidian parasites. Current mt genetic loci used for PCR detection/diagnostics in the genus Eimeria include 805-bp cox1 and 954-bp cox3 [23, 53], and both genetic loci have been recently targeted for development of an amplicon library-based NGS pipeline analysis for diagnosis [13, 23]. From the analysis in this study, however, compared to the cox1 and cox3 genes, it appeared that the cytb gene could also be suitable as a genetic marker for diagnosis and identification between E. maxima/E. tenella and other congeneric species because of its higher variability (Fig. 3c and g). Furthermore, when the sliding window analysis was performed across all chicken Eimeria species, the PCG regions with pronounced peaks and troughs seemed to be more significant than other portions of the mitogenomes (Fig. 3h). Among these PCG regions, cox1 was determined to be the most conserved while cox3 was deemed to be the least conserved, supporting the aforementioned by sequence divergence analysis.

Fig. 3
figure 3

Sliding window analysis of mitogenomes for the nucleotide diversity (Pi) among chicken Eimeria. X-axis: position of the midpoint of a window, Y-axis: nucleotide diversity (Pi) of each window (window size: 200 bp; step size: 20 bp). PCG gene boundaries are indicated above the graph. Colors from gray (high diversity) to white (low diversity) indicate the different nucleotide diversities. a–g Sliding window analysis of each chicken Eimeria species including isolates sequenced in this study. h Sliding window analyses of all chicken Eimeria species including the 33 isolates sequenced in this study. The Pi value of each PCG is shown near the gene name

In parallel with the sliding window analysis, the FST values obtained from PCGs of these Eimeria spp. according to their geographical sources (China vs. Australia) are shown in Additional file 5: Table S5. FST of 0–0.18 suggested species-specific variation in levels of interbreeding, with a higher level of genetic isolation between Chinese and Australian E. mitis populations. In addition, the ratio of non-synonymous (Ka) and synonymous (Ks) substitutions for each PCG in 33 Eimeria isolates and other congeneric species is shown in Fig. 4a–g. Notably, the Ka, Ks and Ka/Ks tied well with results of sliding window analysis, and the PCGs of the Eimeria species with high Pi values trended to have positive selection sites (Ka/Ks > 1). For example, Ka/Ks ratios of the cox1 gene among E. brunetti and E. necatrix isolates and the cox1 and cox3 genes among E. maxima isolates all were > 1. In contrast, the Ka/Ks ratios of the cytb gene in all isolates were observed nearby zero, suggesting its strong purifying selection. Moreover, when the Ka/Ks was calculated across all Eimeria species, all PCG genes exhibited low ratios (< 1) [54], suggesting that these genes were evolving under negative or purifying selection and, to certain extent, implied the conservation and validity of seven Eimeria species infecting chickens.

Fig. 4
figure 4

Evolutionary rates of chicken Eimeria including the 33 isolates identified in this study. Rate of non-synonymous substitutions (Ka), rate of synonymous substitutions (Ks) and ratio of rate of non-synonymous substitutions to rate of synonymous substitutions (Ka/Ks) are calculated for each PCG. a–g Average Ka, Ks and Ka/Ks ratios of three PCGs of isolates reported and sequenced in this study in each species. h Average Ka, Ks and Ka/Ks ratios of 66 Eimeria isolates including the 33 isolates sequenced in this study

Genetic distances of chicken Eimeria

We calculated the intraspecific and interspecific genetic distances between these 33 isolates and each of ten chicken eimerian species using the concatenated PCGs. As shown in Fig. 5, the K2P model-based genetic distances between Eimeria isolates 1–7 and E. tenella, Eimeria isolates 8–13 and E. necatrix, Eimeria isolates 14–19 and E. maxima, Eimeria isolates 20–23 and E. praecox, Eimeria isolates 24–26 and E. mitis, Eimeria isolates 27–30 and E. brunetti and Eimeria isolates 31–33 and E. acervulina approached zero and suggested their species identity, in agreement with the morphological and molecular identifications. For all species clusters, the intraspecific variations were lower than those calculated on the basis of the 18S gene (0.000–0.002 vs. 0.002–0.013) [5]. Moreover, the interspecific genetic distance between E. tenella and E. necatrix was lower than each compared to the other five Eimeria species, once again suggesting that these two species had the closest relationship among chicken eimerian parasites [5, 6, 25, 46,47,48,49,50, 53]. By contrast, the genetic distance between E. maxima and other congeneric species appeared the farthest (0.138 to 0.176), consistent with the previous finding [53], supporting a distinct relationship between E. maxima and other chicken Eimeria species.

Fig. 5
figure 5

Patterns of K2P distance between 33 isolates and 10 known chicken Eimeria species. The edge of these decagons indicates the K2P distance between the isolate and its corresponding species (0.000). Gray lines indicate the same K2P distance from the center of these decagons. Different color dots show the relative K2P distances between isolates with ten Eimeria species. The abbreviations Eac, Ebr, Ela, Ema, Emi, Ena, Ene, Epr, Ete and Eza represent Eimeria acervulina, E. brunetti, E. lata, E. maxima, E. mitis, E. nagambie, E. necatrix, E. praecox, E. tenella and E. zaria, respectively

Phylogenies of chicken Eimeria

The available mitogenomes of 33 isolates representing all seven known chicken eimerian species in China together with an additional 25 mitogenomes of Australian Eimeria spp. provided us an opportunity to study the evolutionary relationships of each isolate in each species and of each species in the genus Eimeria. As shown in Fig. 6a and b, it was clear that three identical trees (MP/ML/BI) inferred from either the complete mitogenomes or concatenated PCG datasets consistently revealed that the Eimeria isolates 1–7 were grouped with available species of E. tenella, the Eimeria isolates 8–13 were grouped with E. necatrix, the Eimeria isolates 14–19 were grouped with E. maxima, the Eimeria isolates 20–23 were grouped with E. praecox, Eimeria isolates 24–26 were grouped with E. mitis, Eimeria isolates 27–30 were grouped with E. brunetti isolates, and Eimeria isolates 31–33 were grouped with E. acervulina, with high statistical supports (all values ≥ 98 or = 1.00). Furthermore, both phylogenetic trees split chicken eimerian parasites into two clades: one included E. tenella and E. necatrix and another was composed of E. lata, E. maxima, E. nagambie, E. brunetti, E. zaria, E. mitis, E. praecox and E. acervulina. Compared to the stable topology of E. tenella and E. necatrix, the phylogenetic relationships among eight other Eimeria species varied by different datasets used here (Fig. 6a and b). Notably, E. maxima and E. mitis were more closely related to each other than to E. praecox and E. acervulina, in contrast with findings based on the nuclear 18S and internal transcribed spacer (ITS) and mitochondrial cox1 datasets [53, 55,56,57]. Nevertheless, these analyses provided a consistent, robust phylogenetic resolution for the 33 isolates and their congeneric species in the genus Eimeria: a paraphyletic relationship was shared among the 10 chicken Eimeria species, in agreement with our genetic distance study herein and the results of morphological and molecular biology studies [6, 56, 58,59,60].

Fig. 6
figure 6

Phylogenetic relationships of chicken Eimeria including the 33 isolates identified in this study. a Phylogenies were inferred on the basis of the complete mitogenome datasets. b Phylogenies were inferred on the basis of the concatenated nucleotide sequences of three PCGs. c Phylogenies were inferred on the basis of the concatenated nucleotide sequences of two PCGs. d Phylogenies were inferred on the basis of the nucleotide sequences of single PCG. 66 chicken Eimeria and one Isospora speecies (outgroup) were included in the aforementioned phylogenetic analyses using the BI, ML and MP methods. Ten chicken Eimeria species, E. acervulina, E. brunetti, E. lata, E. maxima, E. mitis, E. nagambie, E. necatrix, E. praecox, E. tenella and E. zaria, and their branches are shown in dark blue, blue, pink, red, green, cyan, yellow, orange, purple and flavogreen, respectively; 33 Eimeria isolates sequenced in this study are indicated in bold font (a and b) or branches (c and d). The numbers along the branches indicate bootstrap values/posterior probabilities resulting from different analyses in the order MP/ML/BI

Furthermore, the single and concatenated PCG genes were separately used for phylogenetic analysis to screen out the optimal genetic marker candidates for phylogeny and species identification and differentiation of chicken eimerian parasites from other related species. As shown in Fig. 6c and d, although any single or dual combination of PCGs exhibited various phylogenetic topologies, the positions of the species E. tenella and E. necatrix were steady in these six (cox1, cox3, cytb, cytb + cox1, cytb + cox3 and cox1 + cox3) phylogenetic analyses. Notably, the concatenated cytb and cox1 gene- and single cox1-based phylogenetic analyses shared the same topology as that of the genome-based phylogeny, suggesting that the cox1 gene might be the most appropriate genetic marker and therefore could be used instead of mitogenomes for evolutionary and phylogenetic studies of chicken Eimeria species. Of course, the marker validity of the cox1 gene remains further validated when more additional apicomplexan parasite mitogenomes become available, especially those from avain coccidia, although the cox1 has been widely used as a DNA barcode for species identification and differentiation in Eimeria [53, 61].

Conclusions

In this study, we presented a comprehensive molecular characterization of evolutionary blueprint of chicken coccidia by Illumina sequencing complete mitogenomes of 33 isolates representing all seven known Eimeria species infecting chickens in China. Comparative genomics revealed the low genetic diversity of these Eimeria species and showed three mitochondrial protein genes under purifying selection with cox1 and cox3 genes being the lowest and highest evolutionary rates, respectively. Phylogenies divided these 33 Eimeria isolates into seven subgroups, and each represented one chicken Eimeria species. Furthermore, single and concatenated mitochondrial protein gene-based phylogenies supported the cox1 gene as the genetic marker for evolutionary and phylogenetic studies for avain eimerians. These Eimeria population-level mitogenomic datasets provide an updated understanding of systematic, population genetic and evolutionary biological studies of apicomplexan parasites in poultry and other animals.

Availability of data and materials

Molecular data have been deposited to GenBank with the following accession numbers: OP800493–OP800525.

Abbreviations

18S:

Nuclear ribosomal protein 18

ITS-1:

Internal transcribed spacer 1

ITS-2:

Internal transcribed spacer 2

MP:

Maximum parsimony

ML:

Maximum likelihood

BI:

Bayesian inference

PCG:

Protein-coding gene

tRNA:

Transfer RNA

rRNA:

Ribosomal RNA

LSU:

Large subunit ribosomal RNA

SSU:

Small subunit ribosomal RNA

cox1:

Cytochrome oxidase subunits 1

cox3:

Cytochrome oxidase subunits 3

cytb:

Cytochrome b

FST :

Wright’s fixation index

Ka:

Non-synonymous substitutions

Ks:

Synonymous substitutions

References

  1. Fatoba AJ, Adeleke MA. Diagnosis and control of chicken coccidiosis: a recent update. J Parasit Dis. 2018;42:483–93.

    Article  PubMed  PubMed Central  Google Scholar 

  2. Soutter F, Werling D, Tomley FM, Blake DP. Poultry coccidiosis: design and interpretation of vaccine studies. Front Vet Sci. 2020;7:101.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Chapman HD, Barta JR, Blake D, Gruber A, Jenkins M, Smith NC, et al. A selective review of advances in coccidiosis research. Adv Parasitol. 2013;83:93–171.

    Article  PubMed  Google Scholar 

  4. Cantacessi C, Riddell S, Morris GM, Doran T, Woods WG, OtrantoGasser DRB, et al. Genetic characterization of three unique operational taxonomic units of Eimeria from chickens in Australia based on nuclear spacer ribosomal DNA. Vet Parasitol. 2008;152:226–34.

    Article  CAS  PubMed  Google Scholar 

  5. Clark EL, Macdonald SE, Thenmozhi V, Kundu K, Garg R, Kumar S, et al. Cryptic Eimeria genotypes are common across the southern but not northern hemisphere. Int J Parasitol. 2016;46:537–44.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Blake DP, Vrba V, Xia D, Jatau ID, Spiro S, Nolan MJ, et al. Genetic and biological characterisation of three cryptic Eimeria operational taxonomic units that infect chickens (Gallus gallus domesticus). Int J Parasitol. 2021;51:621–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Andreopoulou M, Chaligiannis I, Sotiraki S, Daugschies A, Bangoura B. Prevalence and molecular detection of Eimeria species in different types of poultry in Greece and associated risk factors. Parasitol Res. 2022;121:2051–63.

    Article  CAS  PubMed  Google Scholar 

  8. Gottardo Balestrin PW, Balestrin E, Santiani F, Biezus G, Moraes JC, da Silva Casa M, et al. Prevalence of Eimeria sp. in broiler poultry houses with positive and negative pressure ventilation systems in southern Brazil. Avian Dis. 2021;65:469–73.

    Article  PubMed  Google Scholar 

  9. Karaer Z, Guven E, Akcay A, Kar S, Nalbantoglu S, Cakmak A. Prevalence of subclinical coccidiosis in broiler farms in Turkey. Trop Anim Health Prod. 2012;44:589–94.

    Article  PubMed  Google Scholar 

  10. Kumar S, Garg R, Ram H, Maurya PS, Banerjee PS. Gastrointestinal parasitic infections in chickens of upper gangetic plains of India with special reference to poultry coccidiosis. J Parasit Dis. 2015;39:22–6.

    Article  PubMed  Google Scholar 

  11. Györke A, Pop L, Cozma V. Prevalence and distribution of Eimeria species in broiler chicken farms of different capacities. Parasite. 2013;20:50.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Debbou-Iouknane N, Benbarek H, Ayad A. Prevalence and aetiology of coccidiosis in broiler chickens in Bejaia province, Algeria. Onderstepoort J Vet Res. 2018;85:e1-6.

    Article  PubMed  Google Scholar 

  13. Hauck R, Carrisosa M, McCrea BA, Dormitorio T, Macklin KS. Evaluation of next-generation amplicon sequencing to identify Eimeria spp. of chickens. Avian Dis. 2019;63:577–83.

    Article  PubMed  Google Scholar 

  14. Huang Y, Ruan X, Li L, Zeng M. Prevalence of Eimeria species in domestic chickens in Anhui province, China. J Parasit Dis. 2017;41:1014–9.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Blake DP, Knox J, Dehaeck B, Huntington B, Rathinam T, Ravipati V, et al. Re-calculating the cost of coccidiosis in chickens. Vet Res. 2020;51:115.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Coker S. Morphological and molecular characterisation of coccidia (Eimeria spp.) in kiwi (Apteryx spp.). Animal Science at Massey University, Palmerston North, New Zealand; 2021.

  17. Geng T, Ye C, Lei Z, Shen B, Fang R, Hu M, et al. Prevalence of Eimeria parasites in the Hubei and Henan provinces of China. Parasitol Res. 2021;120:655–63.

    Article  PubMed  Google Scholar 

  18. Vermeulen ET, Lott MJ, Eldridge MD, Power ML. Evaluation of next generation sequencing for the analysis of Eimeria communities in wildlife. J Microbiol Methods. 2016;124:1–9.

    Article  CAS  PubMed  Google Scholar 

  19. Hinsu AT, Thakkar JR, Koringa PG, Vrba V, Jakhesara SJ, Psifidi A, et al. Illumina next generation sequencing for the analysis of Eimeria populations in commercial broilers and indigenous chickens. Front Vet Sci. 2018;30:176.

    Article  Google Scholar 

  20. Barr CM, Neiman M, Taylor DR. Inheritance and recombination of mitochondrial genomes in plants, fungi and animals. New Phytol. 2005;168:39–50.

    Article  CAS  PubMed  Google Scholar 

  21. Lin CP, Danforth BN. How do insect nuclear and mitochondrial gene substitution patterns differ? Insights from Bayesian analyses of combined datasets. Mol Phylogenet Evol. 2004;30:686–702.

    Article  CAS  PubMed  Google Scholar 

  22. Hao W, Richardson AO, Zheng Y, Palmer JD. Gorgeous mosaic of mitochondrial genes created by horizontal transfer and gene conversion. Proc Natl Acad Sci U S A. 2010;107:21576–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Snyder RP, Guerin MT, Hargis BM, Imai R, Kruth PS, Page G, et al. Exploiting digital droplet PCR and next generation sequencing technologies to determine the relative abundance of individual Eimeria species in a DNA sample. Vet Parasitol. 2021;296:109443.

    Article  CAS  PubMed  Google Scholar 

  24. Rokas A, Williams BL, King N, Carroll SB. Genome-scale approaches to resolving incongruence in molecular phylogenies. Nature. 2003;425:798–804.

    Article  CAS  PubMed  Google Scholar 

  25. Morgan JAT, Godwin RM. Mitochondrial genomes of Australian chicken Eimeria support the presence of ten species with low genetic diversity among strains. Vet Parasitol. 2017;243:58–66.

    Article  CAS  PubMed  Google Scholar 

  26. Goan C, Walker F. Poultry litter sampling and testing. In: The university of Tennessee, agricultural extension service. 2009. https://extension.tennessee.edu/publications/Documents/SP563.pdf. Accessed Feb 2018.

  27. Castañón CAB, Fraga JS, Fernandez S, Gruber A, da Costa FL. Biological shape characterization for automatic image recognition and diagnosis of protozoan parasites of the genus Eimeria. Pattern Recognit. 2007;40:1899–910.

    Article  Google Scholar 

  28. Haug A, Gjevre AG, Thebo P, Mattsson JG, Kaldhusdal M. Coccidial infections in commercial broilers: epidemiological aspects and comparison of Eimeria species identification by morphometric and polymerase chain reaction techniques. Avian Pathol. 2008;37:161–70.

    Article  CAS  PubMed  Google Scholar 

  29. Schwarz RS, Jenkins MC, Klopp S, Miska KB. Genomic analysis of Eimeria spp. populations in relation to performance levels of broiler chicken farms in Arkansas and North Carolina. J Parasitol. 2009;95:871–80.

    Article  CAS  PubMed  Google Scholar 

  30. Hahn C, Bachmann L, Chevreux B. Reconstructing mitochondrial genomes directly from genomic next-generation sequencing reads—a baiting and iterative mapping approach. Nucleic Acids Res. 2013;41:e129.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Bernt M, Donath A, Jühling F, Externbrink F, Florentz C, Fritzsch G, et al. MITOS: improved de novo metazoan mitochondrial genome annotation. Mol Phylogenet Evol. 2013;69:313–9.

    Article  PubMed  Google Scholar 

  32. Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, et al. Geneious basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics. 2012;28:1647–9.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Zhang D, Gao F, Jakovlić I, Zou H, Zhang J, Li WX, et al. PhyloSuite: an integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol Ecol Resour. 2020;20:348–55.

    Article  PubMed  Google Scholar 

  34. Perna NT, Kocher TD. Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. J Mol Evol. 1995;41:353–8.

    Article  CAS  PubMed  Google Scholar 

  35. Burland TG. DNASTAR’s lasergene sequence analysis software. Methods Mol Biol. 2000;132:71–91.

    CAS  PubMed  Google Scholar 

  36. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35:1547–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Rozas J, Ferrer-Mata A, Sánchez-DelBarrio JC, Guirao-Rico S, Librado P, Ramos-Onsins SE, et al. DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol Biol Evol. 2017;34:3299–302.

    Article  CAS  PubMed  Google Scholar 

  38. Zhang Z, Li J, Zhao XQ, Wang J, Wong GK, Yu J. KaKs_Calculator: calculating Ka and Ks through model selection and model averaging. Genom Proteom Bioinf. 2006;4:259–63.

    Article  CAS  Google Scholar 

  39. Srivathsan A, Meier R. On the inappropriate use of Kimura-2-parameter (K2P) divergences in the DNA-barcoding literature. Cladistics. 2012;28:190–4.

    Article  PubMed  Google Scholar 

  40. Katoh K, Misawa K, Kuma K, Miyata T. MAFFT: a novel method for rapid multiple sequence alignment based on fast fourier transform. Nucleic Acids Res. 2002;30:3059–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Goloboff PA, Catalano SA, Torres A. Parsimony analysis of phylogenomic datasets (II): evaluation of PAUP*. MEGA and MPBoot Cladistics. 2022;38:126–46.

    Article  CAS  PubMed  Google Scholar 

  42. Castresana J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol. 2000;17:540–52.

    Article  CAS  PubMed  Google Scholar 

  43. Guindon S, Gascuel O. A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst Biol. 2003;52:696–704.

    Article  PubMed  Google Scholar 

  44. Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods. 2017;14:587–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Höhna S, et al. MrBayes 3.2: efficient bayesian phylogenetic inference and model choice across a large model space. Syst Biol. 2012;61:539–42.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Ogedengbe ME, Hafeez MA, Barta JR. Sequencing the complete mitochondrial genome of Eimeria mitis strain USDA 50 (Apicomplexa: Eimeriidae) suggests conserved start positions for mtCOI- and mtCOIII-coding regions. Parasitol Res. 2013;112:4129–36.

    Article  CAS  PubMed  Google Scholar 

  47. Liu G, Li Q, Wang C, Xu C. The complete mitochondrial genome of Eimeria anseris from the wintering greater white-fronted goose in Shengjin Lake, China, and phylogenetic relationships among Eimeria species. Parasitol Res. 2019;118:1299–306.

    Article  PubMed  Google Scholar 

  48. Hafeez MA, Vrba V, Barta JR. The complete mitochondrial genome sequence of Eimeria innocua (Eimeriidae, Coccidia, Apicomplexa). Mitochondrial DNA A DNA Mapp Seq Anal. 2016;27:2805–6.

    CAS  PubMed  Google Scholar 

  49. Liu GH, Hou J, Weng YB, Song HQ, Li S, Yuan ZG, et al. The complete mitochondrial genome sequence of Eimeria mitis (Apicomplexa: Coccidia). Mitochondrial DNA. 2012;23:341–3.

    Article  CAS  PubMed  Google Scholar 

  50. Lin RQ, Qiu LL, Liu GH, Wu XY, Weng YB, Xie WQ, et al. Characterization of the complete mitochondrial genomes of five Eimeria species from domestic chickens. Gene. 2011;480:28–33.

    Article  CAS  PubMed  Google Scholar 

  51. Perkins SL. Molecular systematics of the three mitochondrial protein-coding genes of malaria parasites: corroborative and new evidence for the origins of human malaria. DNA Seq. 2008;19:471–8.

    Article  CAS  Google Scholar 

  52. Hikosaka K, Watanabe YI, Tsuji N, Kita K, Kishine H, Arisue N, et al. Divergence of the mitochondrial genome structure in the apicomplexan parasites, Babesia and Theileria. Mol Biol Evol. 2010;27:1107–16.

    Article  CAS  PubMed  Google Scholar 

  53. Ogedengbe JD, Hanner RH, Barta JR. DNA barcoding identifies Eimeria species and contributes to the phylogenetics of coccidian parasites (Eimeriorina, Apicomplexa, Alveolata). Int J Parasitol. 2011;41:843–50.

    Article  CAS  PubMed  Google Scholar 

  54. Hurst LD. The Ka/Ks ratio: diagnosing the form of sequence evolution. Trends Genet. 2002;18:486.

    Article  PubMed  Google Scholar 

  55. Vrba V, Pakandl M. Host specificity of turkey and chicken Eimeria: controlled cross-transmission studies and a phylogenetic view. Vet Parasitol. 2015;208:118–24.

    Article  PubMed  Google Scholar 

  56. Alam MZ, Dey AR, Rony SA, Parvin S, Akter S. Phylogenetic analysis of Eimeria tenella isolated from the litter of different chicken farms in Mymensingh, Bangladesh. Vet Med Sci. 2022;8:1563–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Fornace KM, Clark EL, Macdonald SE, Namangala B, Karimuribo E, Awuni JA, et al. Occurrence of Eimeria species parasites on small-scale commercial chicken farms in Africa and indication of economic profitability. PLoS ONE. 2013;8:e84254.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Barta JR, Martin DS, Liberator PA, Dashkevicz M, Anderson JW, Feighner SD, et al. Phylogenetic relationships among eight Eimeria species infecting domestic fowl inferred using complete small subunit ribosomal DNA sequences. J Parasitol. 1997;83:262–71.

    Article  CAS  PubMed  Google Scholar 

  59. Miska KB, Schwarz RS, Jenkins MC, Rathinam T, Chapman HD. Molecular characterization and phylogenetic analysis of Eimeria from turkeys and gamebirds: implications for evolutionary relationships in Galliform birds. J Parasitol. 2010;96:982–6.

    Article  CAS  PubMed  Google Scholar 

  60. Hafeez MA, Sattar A, Khalid K, Khalid AR, Mahmood MS, Aleem MT, et al. Molecular and morphological characterization of Eimeria crandallis isolated from deer (Cervidae) in different captive animals. Life (Basel). 2022;12:1621.

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Hafeez MA, Shivaramaiah S, Dorsey KM, Ogedengbe ME, El-Sherry S, Whale J, et al. Simultaneous identification and DNA barcoding of six Eimeria species infecting turkeys using PCR primers targeting the mitochondrial cytochrome c oxidase subunit I (mtCOI) locus. Parasitol Res. 2015;114:1761–8.

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

The authors are grateful to Xiaoyan Zhao, Xue He and Xiaofang Tong who helped in the collection of samples.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 32273028).

Author information

Authors and Affiliations

Authors

Contributions

YX, SW and SC, conceptualization; XZ, LW, PZ and ZW, methodology; XZ, LW and YC, software & data curation; LW, PZ, ZY, XG and BJ, validation; XZ, LW, PZ, RH and JX, investigation; XZ, LW, SC, SW and YX, writing-original draft preparation; YX and GY, writing-review and editing; SC, SW and YX, supervision; SC and YX, project administration & funding acquisition. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Shun Chen, Shuangyang Wu or Yue Xie.

Ethics declarations

Ethics approval and consent to participate

This study was conducted in strict accordance with the recommendations in the Guide for the Care and Use of Committee of Sichuan Agricultural University, China (approval no. SYXK 2019-187). All animal procedures used in this study were carried out in accordance with the Guide for the Care and Use of Laboratory Animals (National Research Council) and recommendations of the ARRIVE guidelines (https://www.nc3rs.org.uk/arrive-guidelines).

Consent for publication

Not applicable.

Competing interests

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

Additional file 1: Table S1.

List of primer pairs for PCR amplifications and their positions in the mitogenome of Eimeria isolate 1.

Additional file 2: Table S2.

Summary of the chicken Eimeria mitogenomic information included in this study.

Additional file 3: Table S3.

Gene sizes in the mitogenomes of 66 chicken Eimeria parasites.

Additional file 4: Table S4.

Pairwise genetic divergences of 33 chicken Eimeria parasites based on mitogenome datasets.

Additional file 5: Table S5.

Wrights Fixation Index (FST) values of Chinese and Australian chicken Eimeria species based on their mitochondrial concatenate PCGs.

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

Zhou, X., Wang, L., Zhu, P. et al. Comprehensive molecular characterization of complete mitogenome assemblies of 33 Eimeria isolates infecting domestic chickens. Parasites Vectors 16, 109 (2023). https://doi.org/10.1186/s13071-023-05712-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13071-023-05712-5

Keywords