Abstract
Hybridogenesis is a rare reproductive mode found in interspecific hybrids that involves discarding one parental genome during gametogenesis and clonally transmitting the other, with the former renewed by backcrossing. It may in theory mediate species range expansion, since such hybrids can also cross and have pure-species offspring for establishing new populations and are often widespread probably due to hybrid vigor. Being inspired by a magnitude difference between estimates of individual heterozygosity, we collected genome-wide data and evaluated predictions from hybridogenesis of hybrids between the wideranging Asian spiny frog Quasipaa boulengeri and a narrow endemic, Q. robertingeri. We rejected other atypical modes of reproduction such as parthenogenesis or androgenesis and provided the first evidence consistent with a hypothesis that hybridogenesis mediated the occurrence of individuals far from the species range (~500 km). Despite contributing to early evidence of the only distinct hybridogenetic complex inferred in ten years, individual heterozygosity has likely been an often overlooked variable. The accumulating genome-wide data may serve as a resource for searching for individual heterozygosity clues of atypical reproduction of interspecific hybrid origin. The findings in spiny frogs make a case that genome-wide data provide crucial evidence for updating our knowledge on reproductive-mode evolution of taxa.
1. Introduction
The recent widespread availability of high-throughput sequencing facilitates reliable comparison of individual heterozygosity that requires a large number of loci (Nei, 1978; Mitton & Pierce, 1980). Extraordinarily high individual heterozygosity often reflects a heterogeneous genetic composition (e.g., Green, 1984; Muniz et al., 2020; Zhang et al., 2021). It can serve as a simple genetic clue to possible hybridization and subsequent processes, especially to hybridogenesis that is also referred to as hemiclonal reproduction (Hotz et al., 1999). This mode has been found in interspecific hybrids and involves the elimination of one parental genome during gametogenesis and clonal transmission of the left one, with the former renewed by backcrossing (Dedukh et al., 2017; Lavanchy & Schwander, 2019). As such, a hybridogenetic system is featured by elevated heterozygosities of F1-hybrid types compared to heterozygosities of the other, pure-species individuals. Partly inspired by heterozygosity patterns, classical, non-social hybridogenesis has been identified in six taxa (Bogart et al., 2007; Lafond et al., 2019; Lavanchy & Schwander, 2019; Suzuki et al., 2020; Majtanova et al., 2021). It occurs in hybrids between the live-bearing fish Poeciliopsis monacha and several congeners in North America (Miller & Schultz, 1959; Schultz, 1969; Mateos & Vrijenhoek, 2005). The other three fish taxa include an Iberian minnow complex Squalius alburnoides (Carmona et al., 1997; Collares-Pereira & Coelho, 2010; Matos et al., 2019), five Hypseleotris gudgeon species in Australia (Bertozzi et al., 2000; Schmidt et al., 2011; Thacker et al., 2022), and three Hexagrammos greenling species in coasts of southern Hokkaido and adjacent areas (Kimura-Kawaguchi et al., 2014; Munehara et al., 2016). It has also been found in hybrids between the European marsh frog Pelophylax ridibundus and two congeners (Berger, 1967; Tunner & Heppich-Tunner, 1991; Dufresnes & Mazepa, 2020; Dufresnes et al., 2024) and between stick insects Bacillus rossius and B. grandii in Sicily (Bullini & Nascetti, 1990; Mantovani et al., 1996; Scavariello et al., 2017; Lavanchy et al., 2024). Although one may share the view that these divergent animals represent a small fraction of hybridogenetic systems (Lavanchy & Schwander, 2019), there have been no new taxa reported for ten years in the era of high-throughput sequencing.
Possibly related to this gap, further evidence than individual heterozygosities is needed to confirm hybridogenesis (Majtánová et al., 2021) or to infer and distinguish it from other systems where hybrids are F1 types. The latter include hybrid sterility and parthenogenesis, gynogenesis, kleptogenesis, and androgenesis of interspecific hybrid origin. In parthenogenesis, the embryo develops from an unfertilized egg. Gynogenesis also transmits the female genome only but requires sperm from a sympatric sexual species to trigger embryogenesis. In kleptogenesis, the sperm acquired/stolen from a co-occurring sexual species usually serve to trigger embryogenesis only. Occasionally, the sperm’s nuclear genome can be incorporated and contribute to the gene pool of an asexual population (Fujita et al., 2020). In androgenesis, the nuclear genome is contributed by the male, while eggs lacking a nucleus are produced by females, or the female genome is eliminated after fertilization (Schwander & Oldroyd, 2016). Genome-wide evidence of parental types and a hybrid type characterized by elevated heterozygosity, ~50/50 ancestry proportions, and a large number of fixed heterozygous loci supports but does not distinguish among these systems. Reflecting the genetic divergence between species, considerable fixed heterozygosity is expected for the F1-hybrid types of hybridogenesis and other modes lacking interspecific recombination, such as ameiotic parthenogenesis and ameiotic androgenesis (e.g., Vrijenhoek et al., 1978; Mantovani & Scali, 1993; Nascetti et al., 2003; Scali et al., 2003; Schwander & Oldroyd, 2016; Flynn et al., 2017; Unmack et al., 2019; Wachi et al., 2021).
On the other hand, several types of evidence may be useful to distinguish hybridogenesis from the others. First, if an F1-hybrid type is widespread, hybrid sterility and modes assuming many local hybrid origins can be rejected by a lack of coexistence of hybrids with both parental species. Second, in lineages exhibiting parthenogenesis, gynogenesis, or kleptogenesis, males are very rare if not absent (Avise, 2008; Tilquin & Kokko, 2016; Fujita et al., 2020), which is not necessarily the case for hybridogenetic lineages (e.g., Schmidt et al., 2011). Third, when a single or few local hybrid origins are assumed, a prevalent local, nuclear haplotype sharing between F1-hybrid and parental types is not expected for androgenesis, parthenogenesis, gynogenesis, and kleptogenesis. In these modes by definition a co-distributed parental species has no or only occasional contribution to the nuclear genome (Lampert & Schartl, 2010; Lehtonen et al., 2013; Morgado-Santos et al., 2017). Similarly, assuming a single or few local hybrid origins, a widespread high mitochondrial DNA similarity between co-distributed F1-hybrid and parental types is not expected for parthenogenesis, gynogenesis, and kleptogenesis (Schlupp, 2005; Bi & Bogart, 2010; Schmidt et al., 2011). A gynogenetic or kleptogenetic lineage is expected to exhibit matrilineal mitochondrial DNA variation/haplotypes different from those of the sperm donor populations (Bogart et al., 2009). In a hybridogenetic system, hybridogens share nuclear and possibly mitochondrial variation with a co-occurring parental species through backcrossing (Doležálková-Kaštánková et al., 2021). Last, the non-hybridogenetic modes, if not combined with for instance a meiotic hybridogenesis-like process that forms homologous haploid gametes (e.g., Arai, 2023; Lafond & Angers, 2024), are not expected to have pure-species offspring. Nevertheless it can be accomplished by hybridogens with eggs and sperm carrying conspecific genomes as found in Pelophylax frogs (Vorburger, 2001a, b; Dufresnes & Mazepa, 2020). For clarity, a summary of such different signatures of hybridogenesis and other systems is provided in Table 1.

Several types of different signatures of reproductive systems where hybrids are F1 types.
If an identified hybridogenetic system does have pure-species offspring of hybridogens, it may be suitable to address an unexplored possibility: disjunct range expansion of a narrowly endemic species mediated by hybridogenesis. An endemic may be restricted by habitat requirements and/or competition from a widespread relative (e.g., McGlone et al., 2001; Matesanz et al., 2009). By forming a hybridogenetic complex with the latter and taking advantage of the expected heterosis (Bulger & Schultz, 1979; Hotz et al., 1999), the former may establish a remote population from pure-species offspring of hybridogens given ecological opportunity. This can be investigated using a hybridogenetic complex ideally between an endemic with disjunct distribution and a widespread species. Such a pattern of distribution, however, has not been observed in the aforementioned six hybridogenetic systems (e.g., Mantovani et al., 1996; Collares-Pereira & Coelho, 2010; Kimura-Kawaguchi et al., 2014; Dufresnes & Mazepa, 2020; Thacker et al., 2022; IUCN, 2024).
In East Asian spiny frogs, Quasipaa robertingeri has a narrow known distribution nested within the range of its widespread close relative Q. boulengeri (Che et al., 2009; Fei, 2020; Yan et al., 2021; Frost, 2023). They inhabit subtropical mountain streams and their banks, with Q. robertingeri recorded only from a small area at the southern edge of the Sichuan Basin (Fig. 1) where sympatric populations of them have been found (Wang et al., 2021). Adults of the two species are considered to be morphologically similar. The tadpole of Q. robertingeri can be readily distinguished from those of Q. boulengeri and closely related Q. shini, Q. spinosa, Q. exilispinosa, and Q. verrucospinosa by the absence of dark transverse markings on dorsal tail (Wu & Zhao, 1995; Fei et al., 2009; Pyron, 2014; Yan et al., 2021). Similar karyotypes of 2n = 26 with five large and eight small chromosome pairs have been reported in Q. robertingeri and Q. boulengeri (Tan & Wu, 1987; Wu & Zhao, 1995; Qing et al., 2012; Xia et al., 2020). In addition, an interchromosomal reciprocal translocation has been inferred to spread among some Q. boulengeri populations along the western edge of the Sichuan Basin but not detected elsewhere (Wang, 2006; Qing et al., 2012).

Sampling sites.
Symbols indicate genotype compositions of samples. Site number colors correspond to clusters BB1 and BR1 (black), BB2 and BR2 (purple), BB3 and BR3 (blue), BB4 and BR4 (white), and RR (red) in Fig. 2.
In a preliminary assessment of the genetic diversity of Q. boulengeri with 32 samples, we have observed a bimodal heterozygosity pattern that cannot be explained by reciprocal translocation. Reciprocal translocation may suppress meiotic recombination of the chromosomes involved and eventually result in high levels of heterozygosity (McKim et al., 1988; Haigis & Dove, 2003). Single nucleotide polymorphisms (SNPs) of these samples were obtained from restriction site-associated DNA sequencing (RAD-seq) data. Some estimates of individual heterozygosity were roughly an order of magnitude higher than the others, and this pattern remained unchanged when analyzing subsets of loci presumably on different chromosomes individually. Moreover, for each of the few loci further examined, a high-heterozygosity sample appeared to have one Q. boulengeri haplotype and one outgroup, Q. robertingeri haplotype. These results can be explained by that the high-heterozygosity samples are F1 type hybrids between the two species.
In this study, we obtained RAD-seq data from samples across the distribution of both species and conducted heterozygosity, ancestry proportion, principal coordinate plot, individual clustering, population tree, haplotype network, and mitochondrial gene tree analyses. We first provided evidence supporting a system where hybrids were F1 types. By integrating our results and the literature, we rejected other atypical reproductive modes and accepted hybridogenesis that could explain the occurrence of pure individuals of Q. robertingeri far from its narrow, presumed original range.
2. Materials and methods
2.1. Sampling
A total of 107 individuals from 53 representative localities throughout the range of Quasipaa boulengeri and Q. robertingeri were included as ingroup members (Fig. 1). Among them, there were 58 sexed samples, 27 males and 31 females. The others were mostly tadpoles. Among the sexed samples, 31 individuals had been previously karyotyped, including some reciprocal translocation carriers (Qing et al., 2012; Xia et al., 2020). Based on the current phylogenetic hypothesis for Quasipaa (Yan et al., 2021), the species Q. exilispinosa, with four individuals from three sites, was used as outgroup when necessary. Altogether 111 samples were subjected to RAD-seq, including the aforementioned 32 samples for a preliminary analysis. Details of samples are provided in Table S1.
2.2. RAD-seq and data processing
Liver, muscle, or tail clips preserved in 95% ethanol were sent to Novogene (Beijing, China) for genomic DNA extraction, library construction, and sequencing. The library was prepared following the SLAF-seq approach (Sun et al., 2013) with a target insert size range of 215–240 bp. The restriction enzyme MseI was used in the restriction-ligation reaction, and a subset of the product was excluded by being cut by the restriction enzyme HaeIII. Enriched libraries were sequenced on an Illumina Novaseq 6000 platform in PE150 mode.
The quality of raw reads was checked using FastQC version 0.11.9 (Andrews, 2010). Exclusion of reads containing adapter sequences was conducted by Novogene. Further data cleaning was performed using the process_radtags program in Stacks version 2.60 (Catchen et al., 2013), removing reads with an average Phred33 score of < 10 within a 22-bp sliding window or an uncalled base.
We conducted a de novo assembly of loci using ipyrad version 0.9.87 (Eaton & Overcast, 2020), which applied reverse-complement clustering to our single-enzyme RAD-seq reads. To complete the assembly in a reasonable time, the clean reads were trimmed and filtered to avoid overloading with singlecopy reads according to the ipyrad documentation (https://ipyrad.readthedocs.io/). Single-copy reads can result from sequencing errors that are more frequent toward the 3′ end of reads. The last five bases of each read were trimmed using Trimmomatic version 0.39 (Bolger et al., 2014). Then, for each sample, a file containing only single-copy reads was obtained. This was achieved by sequentially using the dedupe.sh script in BBMap version 38.87 (Bushnell, 2014), the ‘common’ option in SeqKit version 2.3.0 (Shen et al., 2016), and the filterbyname.sh script in BBMap. According to this file, single-copy reads were removed by filterbyname.sh. In the ipyrad pipeline, 3′-end bases with Phred33 scores below 20 were trimmed and three mitochondrial genomes were used as reference for excluding potential mitochondrial reads. They were under GenBank accession numbers KC686711 (Q. boulengeri; Shan et al., 2014), MZ895125 (Q. robertingeri; direct submission by Wang B and Shu G), and MT561179 (Q. exilispinosa; Wu et al., 2020). A minimum depth of 6 for both statistical and majority-rule base calling, a maximum cluster depth of 200 within samples, and an up to 70% occurrence for a shared heterozygous site were allowed. In a preliminary analysis of 22 representative ingroup samples, results obtained with clustering threshold values from 0.74 to 0.95 were compared. This parameter was determined as 0.89 according to the average observed heterozygosity, Pearson coefficient of correlation between missingness and genetic distance, and cumulative variance of the first five principal components (Ilut et al., 2014; Mastretta-Yanes et al., 2015; McCartney-Melstad et al., 2019). The correlation coefficient and cumulative variance were calculated using the vcfMissingness.pl and vcfToPCAvarExplained.R scripts (McCartney-Melstad et al., 2019), respectively. The maximum fraction of heterozygous bases in a consensus sequence was set to 0.11, i.e., one minus the clustering threshold as in the preliminary analysis for possible F1-hybrid types.
The ipyrad SNP output was filtered using VCFtools version 0.1.17 (Danecek et al., 2011) by retaining genotypes with read depth ≥ 6 and SNPs present in ≥ 60% of ingroup members. To minimize linkage disequilibrium, the randSnps.pl script (https://www.biostars.org/p/313701/) was used to extract one random SNP per locus, resulting in the final main dataset. In addition, a dataset containing mainly pure Q. robertingeri samples was generated for population tree reconstruction. Two Q. boulengeri samples from one locality were included for rooting purposes. The ipyrad output was filtered with VCFtools by keeping genotypes with read depth ≥ 6 and SNPs present in all pure Q. robertingeri samples. Then, one SNP was randomly selected for each locus by randSnps.pl.
For fixed heterozygosity and haplotype network analyses, we also used Stacks to perform loci assembly of ingroup members. Stacks provides the observed heterozygosity for each SNP for each sample group and infers haplotype sequences using an algorithm relying on the co-observation of alleles in a read or read pair (Rochette et al., 2019). Currently, ipyrad provides unphased alleles as stated in its documentation. The denovo_map.pl pipeline of Stacks was applied to the clean reads, and a minimum of 3 reads was required to form a stack. Given the read length of 144 bp, the aforementioned ipyrad clustering threshold (0.89) suggested allowing 16 (i.e., 144 − 144 × 0.89) mismatches between stacks and between sample loci. These allowed mismatches were however both reduced to 12 for the reason of computational time based on a preliminary analysis. Loci present in ≤ 60% of samples within a sample group were retained for subsequent analyses.
2.3. Individual heterozygosity, ancestry proportion, and fixed heterozygosity analyses
For each individual, the proportion of heterozygous SNPs was obtained from the main dataset and subsets of it using the --het option in VCFtools. For each SNP in this dataset, the first locus sequence in the gphocs output file from ipyrad was extracted with a custom script gphocs1stSeq2fasta.pl (https://doi.org/10.57760/sciencedb.14822) (will be published on ScienceDB upon final acceptance of the manuscript; data private access link: https://www.scidb.cn/en/s/URnIba) and TBtools version 1.120 (Chen et al., 2020). Using BLAST+ version 2.13.0 (Camacho et al., 2009) with an expectation value threshold of 1e-10, these sequences were compared against the chromosome-level genomes of Q. spinosa (Hu, 2021; Hu et al., 2022) and Nanorana parkeri (Genome Warehouse accession GWHBCKU00000000; Fu et al., 2022). Nanorana is closely related to Quasipaa (Pyron, 2014). Then loci with unique BLAST hits on homologous scaffolds of the two genomes were listed to obtain the proportion of heterozygous SNPs corresponding to individual and all 13 major scaffolds.
Ancestry proportions were inferred using ADMIXTURE version 1.3.0 (Alexander et al., 2009) based on the main dataset, which had been converted into BED input files by PLINK version 1.9 (Chang et al., 2015). For ingroup members, the number of ancestral populations (K) was set to 2. This value could be expected to result in only pure-species types and F1-hybrid types with ancestry proportions close to 50/50 when hybridogenesis occurred (e.g., Kimura-Kawaguchi et al., 2014; Dubey et al., 2019; Lavanchy et al., 2024). To identify possible intraspecific hybrids, K values ranging from 1 to 10 were applied to a subset of samples in separate runs. Then the results from K values with minimum cross-validation errors were compared.
To examine the source of extraordinarily high individual heterozygosity, a simplified 12-individual dataset was extracted from the main dataset. It contained no missing data and three sample types according to ancestry proportions. In each type four samples were selected by having the most loci. Custom scripts were written to find SNPs being heterozygous for one type but showing alternatively fixed alleles, or the same allele, in the other two types.
For fixed heterozygosity analysis, given the ancestry proportion estimates, two Stacks sample groups with the same number of individuals were defined: F1-hybrid types and a randomly selected subset of pure Q. boulengeri types. If a retained Stacks locus was found to contain a SNP with an observed heterozygosity of 1 for a sample group, it was selected. Subsequently, using BLAST+ and an expectation value threshold of 1e-10, its consensus sequence was compared against the aforementioned ipyrad sequences of loci with unique hits on homologous scaffolds.
2.4. Nuclear haplotype networks
Haplotype network construction was conducted using Stacks loci that matched an ipyrad locus with unique-hits on homologous scaffolds and exhibited fixed heterozygosity in F1-hybrid types. For each of the 13 major scaffolds, two loci also present in ≥ 60% of pure-species types were randomly selected for analysis. Sequences with any undetermined bases were excluded. Statistical parsimony network analysis was performed using TCS version 1.21 (Clement et al., 2000) at the 95% confidence level.
2.5. Mitochondrial DNA data and tree
Mitochondrial DNA sequences from three resources were combined to form a single dataset. First, mitochondrial sequences were extracted from the RAD-seq reads cleaned by process_radtags for ingroup members (Barth et al., 2020; Laczkó et al., 2022). MITObim version 1.9.1 (Hahn et al., 2013) was used to assemble mitochondrial sequences, with the aforementioned mitochondrial genomes of Q. boulengeri and Q. robertingeri as references. Reads were mapped to assembled sequences with BWA version 0.7.17 (Li & Durbin, 2009) to obtain the read depth of each site using SAMtools version 1.9 (Li et al., 2009). Sites with read depth < 6 were masked using BEDTools version 2.25.0 (Quinlan & Hall, 2010). Second, four mitochondrial DNA fragments, atp6-cox3, nad4, nad5, and cytb, were PCR amplified and Sanger sequenced (ABI 3730) with primers in Table S2. For a subset of 59 samples, in most cases, all four fragments were sequenced. Lastly, published sequences of the cytb and cox1 genes were retrieved from GenBank for 24 and 35 samples, respectively (Qing et al., 2012; Xia et al., 2016; Cao et al., 2018; Yuan et al., 2018) (Table S1). We also used the published mitochondrial genome of Q. verrucospinosa (KF199147; Zhang et al., 2018) and the aforementioned mitochondrial genome of Q. exilispinosa as outgroups according to current mitochondrial genome tree reconstructions (Zhang et al., 2018). The mitochondrial genome of Q. boulengeri has been found to exhibit tandem duplication and random loss of trnW, trnA, trnN, light-strand origin of replication, trnC, and trnY (Xia et al., 2016). It may also be the case for Q. robertingeri (Zhang et al., 2018). This “WANCY” region was not included. Alignment was performed using ClustalX version 2.1 (Larkin et al., 2007) and checked by eye. Regions with alignment gaps in the ingroup were excluded.
A maximum likelihood tree was inferred from the combined dataset using RAxML version 8.2.12 (Stamatakis, 2014) with 1000 inferences and 2000 non-parametric bootstrap replicates. The rapid hillclimbing algorithm and automatic determination of the initial rearrangement setting were used. The bestfitting combination of data partitions and substitution models was determined using PartitionFinder version 2.1.1 (Lanfear et al., 2017) under the corrected Akaike information criterion. To mitigate potential overparameterization in an analysis of close relatives, only seven data blocks were defined for PartitionFinder evaluation. They were tRNA genes, 12S rRNA gene, 16S rRNA gene, the first, second, and third codon positions of protein-coding genes, and the D-loop region. The GTR+I+G substitution model was not evaluated due to the partial overlapping of the invariable sites parameter I and the among-site rate variation parameter G.
2.6. Principal coordinate plot, neighbor-joining tree, and population tree
To further assess genetic structure, using the main dataset and PLINK, an identity-by-state distance matrix between individuals was calculated for principal coordinates and neighbor-joining cluster analyses. Principal coordinates analysis was performed using GenAlEx version 6.502 (Peakall & Smouse, 2012) for ingroup members. A neighbor-joining tree was constructed with 1000 bootstrap replicates using the Neighbor program in PHYLIP version 3.695 (Felsenstein, 2013). The script VCFbootstrap.pl (Zheng et al., 2020) was used to generate the 1000 bootstrap datasets.
To further examine relationships among pure Q. robertingeri samples and especially between those within and outside of the known species range, a population-tree approach was adopted. A maximum likelihood inference of population splitting and mixing was conducted using SNP allele frequencies and TreeMix version 1.13 (Pickrell & Pritchard, 2012). The allele frequencies for each sampling locality were calculated using PLINK. In separate TreeMix analyses, models allowing zero to six migration events were explored, and the results were compared.
3. Results
3.1. Individual heterozygosity, ancestry proportion, and fixed heterozygosity
The ipyrad genome-wide main dataset contained 45,523 SNPs, 8740−42,842 (mean = 30,600; 67.2%) for ingroup members and 2679−13,244 (8488) for outgroup individuals. Some individuals had much higher heterozygosities than the others, and the two forms occurred together and separately in the samples. The low- and high-heterozygosity samples were inferred as pure-species types and F1-hybrid types with approximately 50/50 ancestry proportions, respectively (Fig. 2). From 21 sites mainly around the Sichuan Basin, 40 F1 types showed similar proportions of heterozygous loci with a mean of 0.167 and SD 0.004. For the other 71 samples including outgroups, a mean value of 0.029 and SD 0.015 were estimated, with a range from 0.009 found in site 30 to 0.081 for sample 28B (Figs. 1 and 2). Sexed F1 types came from 13 sites, 18 females from 9 sites and 10 males from 7 sites. The interchromosomal reciprocal translocation carriers were F1 types, and F1-type tadpoles had distinct, weak, or no transverse dark markings on dorsal tail (Table S1). For the 55 pure Q. boulengeri types, the smallest cross-validation errors were obtained with ADMIXTURE K priors 2 to 4. Sample 28B was the only one that appeared to be an intraspecific hybrid in all three cases (Fig. S1). The samples showing the lowest heterozygosities, 30B and 30D, were inferred to be pure Q. robertingeri types, the other ten of which were found in the known species range (Fig. 1).

Principal coordinate plot (A), heterozygosity pattern (B), ADMIXTURE ancestry proportions (C), and neighbor-joining tree (D) based on the 45,523-SNP main dataset.
Italic numbers on the neighborjoining tree indicate sampling sites in Fig. 1. Numbers beside nodes are bootstrap support values ≥70, and circles on nodes indicate support values of 100. The sample 28B, a putative hybrid of different Quasipaa boulengeri lineages, is marked by an asterisk. In the neighbor-joining tree, the two individuals showing Q. robertingeri genotypes found far from the known species range are highlighted in purple. BB: Q. boulengeri; RR: Q. robertingeri; BR: putative Q. boulengeri × Q. robertingeri hybridogen; Qe: four Q. exilispinosa individuals.
In the main dataset, 20,912 SNPs were extracted from loci with unique BLAST hits on homologous major scaffolds of the Q. spinosa and N. parkeri genomes. These comprised 2379-3467 loci for the five larger scaffolds and 549-1219 loci for the eight smaller ones. Based on these SNPs, the above distribution pattern of heterozygosity among samples remained unchanged for all and individual scaffolds (Table S1). Besides, two datasets containing at least 70% or 80% ingroup members at each locus were generated. Compared with the main dataset, they showed similar heterozygosity, ancestry proportions, principal coordinate plot, and neighbor-joining clustering results. No further analyses were applied to the two datasets. For clarity, a summary of most datasets used in this study is shown in Table 2.

Dataset summaries.
Unique hit: loci with unique BLAST hits on homologous chromosome-level scaffolds of two genomes of the related Quasipaa spinosa and Nanorana parkeri; BB: Q. boulengeri; RR: Q. robertingeri; BR: putative Q. boulengeri × Q. robertingeri hybridogen.
The 100% complete, 12-individual subset of the main dataset contained 25,046 SNPs. A total of 2567 sites were found to be heterozygous in the F1 type and alternatively fixed in the two pure-species types, constituting the majority of the high heterozygosity estimates. The same pattern was observed when only SNPs from loci with unique BLAST hits were considered (Table 3). No sites were found to be heterozygous in the F1 type while showing the same allele in both pure-species types.

Heterozygous SNP numbers of twelve individuals in subsets of the 45,523-SNP main dataset.
The 25,046-site subset contained no missing data and 12,183 sites each corresponding to a locus with unique BLAST hits on homologous scaffolds of two published genomes of Quasipaa spinosa and Nanorana parkeri. ID: individual ID; BB: Q. boulengeri; RR: Q. robertingeri; BR: putative BB x RR hybridogen. SNPs being heterozygous for BR while showing one allele in BB and the other in RR were considered (H) or excluded (H′).
A considerable amount of fixed heterozygosity of the F1 type was detected. A total of 1260 Stacks loci exhibiting fixed heterozygosity among the 40 F1 types (≥ 60% completeness) were found to have a BLAST hit to the aforementioned 20,912 unique-hit ipyrad loci. These comprised 156−210 and 24−66 loci corresponding to the larger and smaller scaffolds, respectively. No such fixed heterozygous loci were detected among the 40 randomly selected pure Q. boulengeri types.
3.2. Nuclear haplotype networks and sharing
For each of the 26 representative Stacks loci for fixed heterozygosity, a single parsimony network was constructed (Figs. 3 and S2). Typically more haplotypes were found for the pure Q. boulengeri type (mean = 6.7) than for the pure Q. robertingeri type (1.6). Haplotypes were rarely shared between the two types, involving only three loci and three haplotypes. On the other hand, an F1 type almost always shared one haplotype with the pure Q. boulengeri type and the other with the pure Q. robertingeri type. In the few exceptional cases between F1 and Q. robertingeri types, a single substitution instead of haplotype sharing was observed. This allowed a ready identification of putative Q. robertingeri haplotypes.

Statistical parsimony network for haplotypes of a single exemplary locus, CLocus_761519 of 144 bp, from individuals of Quasipaa boulengeri, Q. robertingeri, and putative Q. boulengeri × Q. robertingeri hybridogen genotypes.
This locus was heterozygous in putative hybridogens and had unique BLAST hits on the first scaffolds of published Q. spinosa and Nanorana parkeri genomes. Lines represent single substitutions. Filled circles are haplotypes not encountered. The other circles’ sizes are proportional to haplotype frequency. Putative hybridogens’ sample IDs are shown. Samples of Q. robertingeri exhibited haplotypes marked by asterisks.
Compared with the pure Q. robertingeri type, the F1 type usually showed the same or fewer putative Q. robertingeri haplotypes for each locus. For all 26 loci, the former (
Haplotype sharing between F1 and pure-species types in the same locality was common across the 26 loci. For the 21 F1 types co-occurring with pure Q. boulengeri ones at nine sites, it was observed in 94.4% of the 357 cases with available data. For the two F1 types sampled with pure Q. robertingeri ones at site 30, it was found in 42 of 43 cases. Of the 26 loci, 21 had at least five haplotypes found in pure Q. boulengeri ones. Besides, six F1-type-sample only sites had a nearby pure Q. boulengeri sample, i.e., sites 1, 2, 4, 33, 34, and 41 (Fig. 1). When these sites were also considered, local haplotype sharing between F1 and pure Q. boulengeri types was observed in 92.8% of the 347 cases across the 21 loci.
3.3. Mitochondrial DNA data and tree
Mitochondrial sequence fragments were assembled from RAD-seq reads for all ingroup members, from 2472 to 6983 bp per sample with an average of 5099 bp. A total of 179 newly obtained Sanger sequences were deposited under GenBank accessions PP065924–36 and PP083013–178 (Table S1). The combined mitochondrial dataset contained 10,568 sites and was 52.7% complete. For ingroup members, 1514 sites were variable and 1154 sites were potentially parsimony-informative. This dataset is available on ScienceDB (https://doi.org/10.57760/sciencedb.14822) (will be published upon final acceptance of the manuscript; data private access link: https://www.scidb.cn/en/s/URnIba). For the RAxML analysis, a six-partition scheme with a GTR+G model for each partition was selected: tRNA, rRNA, D-loop, and each of the three codon positions.
F1 types exhibited mitochondrial haplotypes clustering together with those of co-occurring and nearby pure Q. boulengeri types on the generally well-supported RAxML tree (Fig. 4) and often shared haplotypes with the latter. In the Q. robertingeri clade Qr, haplotypes from the eastern and western known distribution range were grouped separately. In the other clade Qb, haplotypes clustered into four subclades.

Maximum likelihood population tree of Quasipaa robertingeri with three migration edges (arrows) inferred using TreeMix (A) and RAxML mitochondrial gene tree (B).
The TreeMix tree was based on the 8,535-SNP nuclear dataset of 100% completeness for ingroups. The number below an arrow is migration weight that represents the fraction of ancestry derived from the migration edge. s.e.: average standard error of the sample covariance matrix of allele frequencies. The RAxML tree was based on a 10,568-site dataset of 52.7% completeness and six partitions. Italic numbers beside nodes are bootstrap support values ≥70, and circles on nodes indicate support values of 100 (closed) and 98–99 (open). The characters BB, RR, or BR following the sample ID indicate Q. boulengeri, Q. robertingeri, or putative hybridogen nuclear genotypes, respectively. Lineage names within the Qb clade correspond to cluster names in Fig. 2. Outgroups are not shown for the RAxML tree.
Mitochondrial haplotypes from Q. boulengeri were detected in three pure Q. robertingeri types. Sample 21A had a haplotype closely related to that of a nearby F1 type, the others, at site 30, shared haplotypes with co-occurring F1 or nearby pure Q. boulengeri types. No mitochondrial haplotypes from Q. robertingeri were found in pure Q. boulengeri types.
3.4. Principal coordinate plot, neighbor-joining tree, and population tree
The principal coordinate plot showed three discrete clusters corresponding to the F1 and two pure-species types, with the former located approximately intermediate between the latter two (Fig. 2). Among them, 71.1% of the observed genetic variation was explained by the first two principal coordinates. Much more variation was observed within the F1 or pure Q. boulengeri types than in the pure Q. robertingeri type. Subclusters could be identified in the former two.
The neighbor-joining tree was generally well supported. Two pure-species clusters with bootstrap support of 100 were observed. The two pure Q. robertingeri types from site 30 were grouped with those from the eastern known species range (sites 23–25 in Fig. 1) with strong support. For the F1 type, subclusters BR1, BR2, BR3, and BR4 were identified, corresponding to the four subclusters of the pure Q. boulengeri type, BB1, BB2, BB3, and BB4, respectively (Fig. 2D). When applying these subclusters to the principal coordinate plot, parallel structures were readily detected. Along the second principal coordinate, a BB2+BB3 cluster was located between BB1 and BB4, and BR2+BR3 was located between BR1 and BR4. Besides, BB1–4 and BR1–4 corresponded to the four subclades of the mitochondrial clade Qb (Fig. 4). The full neighbor-joining tree is available as Fig. S3.
The dataset for TreeMix analysis of pure Q. robertingeri types contained 8535 SNPs. Three migration edges between populations from the known species range were found to increase the likelihood significantly (all P < 0.01). With these edges added or not, the same maximum likelihood tree topology was obtained in which site 30 was again nested within the eastern clade (Fig. 4).
4. Discussion
4.1. Hybridogenesis and insular occurrence
In the Quasipaa boulengeri-Q. robertingeri complex, hybridogenesis can be inferred and distinguished from other systems where hybrids are F1 types. Our dataset consisted of two low-heterozygosity parental types and one hybrid type characterized by much higher genome-wide heterozygosity (Fig. 2B), ~50/50 ancestry proportions (Fig. 2C), and a large number of fixed heterozygous loci (6%) throughout the genome. Previous karyotyping of these frogs has identified solely 2n = 26 diploid individuals (e.g., Tan & Wu, 1987; Xia et al., 2020), which when genotyped appear as either F1 or parental types. With F1 type hybrids widely detected (Fig. 1), the known natural range of Q. robertingeri is narrow (e.g., Qing et al., 2012; Yan et al., 2013; Fei, 2020; Wang et al., 2021). Hybrid sterility, or any mode requiring many local hybrid origins, can be ruled out by such a lack of coexistence of hybrids with both parental species (Table 1). In the present study, 10 of the 27 sexed F1 type adults were males from 7 sites (Table S1). This is not expected for parthenogenesis, gynogenesis, or kleptogenesis. In lineages exhibiting such a mode, males are absent or very rare (Avise, 2008; Tilquin & Kokko, 2016; Fujita et al., 2020). For 17 sites across the F1 type range, local nuclear haplotype sharing between F1 and pure Q. boulengeri types occurred at a rate of 92.8% in 21 representative loci. This is unexpected for androgenesis, parthenogenesis, gynogenesis, and kleptogenesis when assuming a single or few local hybrid origins. In these modes a codistributed parental species has only occasional (kleptogenesis) or no contribution to the nuclear genome (Lampert & Schartl, 2010; Lehtonen et al., 2013; Morgado-Santos et al., 2017). Furthermore, the widespread high mitochondrial DNA similarity between co-distributed F1 and Q. boulengeri types (Fig. 4B) is not expected for parthenogenesis, gynogenesis, and kleptogenesis of a single or few local hybrid origins (Bogart et al., 2009; Bi & Bogart, 2010; Schmidt et al., 2011). Last, the two pure Q. robertingeri types found with F1 types at site 30 can hardly be explained by even a combination of non-hybridogenetic processes. With divergent non-conspecific mitochondrial genomes (Plotner et al., 2008; Liu et al., 2010), these two samples were nested deeply among those from the narrow known range approximately 500 kilometers away based on genome-wide nuclear data (Figs. 2D and 4). It is unlikely that, through androgenesis, they have been produced by parthenogenetic hybrid females and local relict Q. robertingeri males (Tinti & Scali, 1996). They may also in theory have arisen from androgenesis with homologous haploid sperm formed by hybrids through a meiotic hybridogenesis-like process (Arai, 2023; Lafond & Angers, 2024). However, their low but apparent and similar chromosome-level heterozygosities indicate other infrequent events, polyspermy involving different males (Talevi, 1989; Schwander & Oldroyd, 2016; Hater et al., 2020; Iwao & Ueno, 2022). In contrast, all the above findings are consistent with a hybridogenetic system in which hybridogens can share nuclear and mitochondrial genetic variation with local parental species and can produce heterozygous pure-species offspring (Vorburger, 2001a, b; Dufresnes & Mazepa, 2020; Doležálková-Kaštánková et al., 2021).
Along with the clustering of pure Q. robertingeri types, two lines of evidence also support an “out of the known range” origin for the widespread Q. robertingeri nuclear genomes. First, the mitochondrial lineage corresponding to this species is endemic to the narrow known range, and it shows a significant east-west division that suggests long-term persistence (Figs. 1 and 4B) (Yan et al., 2013; Wang et al., 2021). Second, for nuclear loci, the variation detected within the known range is greater than and overlaps substantially with that found outside it. For the 26 representative loci, on average 7 pure-species and 28 F1/pure-species types were respectively sampled within and outside the known range. However, 23.5% more putative Q. robertingeri haplotypes were detected in the former, with 88.2% such haplotypes observed in the latter shared by the former. Based on the 45,523-SNP data, parallel geographic structures were inferred for pure Q. boulengeri and F1 types (Figs. 1 and 2A). This also agrees with the “out of the known range” scenario. Because nuclear variation in Q. boulengeri was much higher than that of Q. robertingeri from the known range, among hybridogens the former could contribute dominantly in characterizing geographic structure.
In addition, it is unlikely that accidental transport of tadpoles (along the Yangtze River) or human-mediated translocation from the known range has caused the insular occurrence of Q. robertingeri nuclear genomes at site 30 (Fig. 1). In both cases, the F1 types of this site are expected to have a local Q. boulengeri haploid nuclear genome and hence belong to subcluster BR1. However, they were nested in subcluster BR4 and were most closely related to the F1 type from an intermediate site (28) between the known range and their locality (Fig. S3). Moreover, three of the four samples from site 30 also had mitochondrial haplotypes most closely related to the one found in site 28 (Fig. 4B).
To our knowledge, this study provides the first evidence consistent with a hypothesis that hybridogenesis mediates the occurrence of individuals far from the continuous distribution of conspecifics. Such an occurrence obviously does not guarantee the successful establishment of a Q. robertingeri population, but the possibility deserves further exploration in addition to cytogenetic examination of hybridogenesis. The reason is that if this kind of reconstituted population does exist, then the finding will introduce a new mode of range expansion and extend our understanding of the interaction between narrow-ranged and widespread relatives.
4.2. Individual heterozygosity and features of the spiny frog system
Atypical reproductive systems of hybrid origin are unusual and useful research resources (Laskowski et al., 2019). For instance, hybridogenesis suggests an approach to preserve a species in non-optimal conditions by taking advantage of a relative and of hybrid vigor (Dubey & Dufresnes, 2017). This study constitutes another example of heterozygosity patterns inspiring a test of atypical reproduction, with the discovery of a putative hybridogenetic complex unrelated to any known one for the first time since 2014 (Kimura-Kawaguchi et al., 2014; Lavanchy & Schwander, 2019; Dufresnes et al., 2024). From a dataset of genome-wide loci, individual heterozygosity estimates are among the first statistics that one can obtain. If some estimates are much higher than the others which is the case here, the pattern can serve as a clue to the presence of unusual reproductive modes (e.g., Jaron et al., 2021; Lavanchy et al., 2024). Such datasets are common in the era of high-throughput sequencing. However, a simple Google Scholar search has returned only 1850 or 812 articles dated between 2014 and June 3, 2025 with the phrase “individual heterozygosity/heterozygosities” or “proportion/proportions of heterozygous loci/sites”, respectively. It implies that this variable has not been reported in most cases, making the accumulating genome-wide data a resource for searching for individual heterozygosity clues of atypical reproduction of interspecific hybrid origin.
Compared with documented hybridogenetic systems, Q. boulengeri and Q. robertingeri occur in a relatively poorly studied region from a natural history perspective, East Asian mainland. The former have been reported from North America, Europe, Australia, and the Japanese islands (Suzuki et al., 2020; Majtánová et al., 2021). Research efforts are not equally distributed across taxa. From under-studied taxa, atypical reproduction may even be found as a by-product of efforts to understand the study system.
As revealed by conventional staining, the Q. boulengeri x Q. robertingeri system shows structurally abnormal, non-Y/W chromosomes which are regionally widespread. The fixation or partial fixation of such chromosomal mutations is generally rare probably due to deleterious effects and meiotic pairing abnormalities (Wallace et al., 1999; Bidau & Marti, 2010; Mezzasalma et al., 2017, 2021). In hybridogenesis, clonal genomes, which do not face meiotic pairing problems and recombination, are prone to accumulation of deleterious mutations and can exhibit diagnostic chromosomal rearrangements (Suzuki et al., 2017, 2020; Lavanchy & Schwander, 2019; Dufresnes & Mazepa, 2020). In the spiny frog system, female and male carriers of an interchromosomal reciprocal translocation (karyotype IRT) have been found in some populations at the western edge of the Sichuan Basin (Fig. 1). Furthermore, there is a tight linkage between the two involved chromosomes, with other co-occurring abnormal karyotypes rarely observed (Qing et al., 2012). In this region, our samples with normal karyotype can be either pure Q. boulengeri (n = 10) or putative hybridogens (n = 4). As predicted by a hypothesis that the linkage is maintained by clonal transmission of the Q. robertingeri haploid genome, the 6 samples of karyotype IRT all belong to the latter (chi-square P < 0.01). The rest few samples of rare abnormal karyotypes are also putative hybridogens showing chromosome-level heterozygosity estimates of similarly large values, implying a complicated scenario with for instance possible subsequent rearrangements. When non-sex-specific abnormal chromosomes are encountered, hybridogenesis may serve as a hypothesis to be tested.
Quasipaa boulengeri and Q. robertingeri show significant mitochondrial divergence comparable to those between parental species of known hybridogenetic systems. These parental species are expected to be similar enough to produce fertile offspring. They are also expected to attain a certain divergence before hybridization leads to a transition to hybridogenesis rather than to genetic admixture (Dufresnes & Mazepa, 2020; Barley et al., 2022). Currently, genome-wide nuclear data for a comparison across known hybridogenetic systems are not available. However, a mitochondrial cytb partial sequence of 948 bp is accessible for most of these parental species in GenBank. By arbitrarily using one such sequence to represent a Pelophylax frog species (JN627421, 22, MF667576; Hofman et al., 2012; Dufresnes et al., 2017), one can obtain uncorrected p-distances of 0.141 and 0.133. They are for the two parental species pairs P. ridibundus–P lessonae and P. ridibundus–P perezi (Dufresnes & Mazepa, 2020), respectively. The corresponding value is 0.113 between Q. boulengeri (KC686711) and Q. robertingeri (MZ895125). For hybridogenetic fishes, such p-distances can be estimated as 0.100 and 0.082 in Hexagrammos (AB290742, 48, 69; Kimura et al., 2007), 0.129–0.133 in Poeciliopsis (AF412139, 51, MK860197, 98, KX229692; Mateos et al., 2002, 2019; Jeon et al., 2016), and 0.098–0.103 in Hypseleotris (DQ468188, 236, 250, 296; Thacker et al., 2007). In Squalius, a mitochondrial lineage corresponding to one of the two parental species has not been found (Matos et al., 2019). In the other, Bacillus stick insect hybridogenetic system, a similar level of mitochondrial divergence can also be observed. An uncorrected p-distance of 0.074 is estimated for a 675-bp fragment of the cox2 gene (AF148299, 315; Mantovani et al., 2001), which evolves at a rate close to that of the cytb gene (e.g., Hofman et al., 2012). These estimates are higher than many shallow interspecific distances (e.g., Young et al., 2013; Tarvin et al., 2017), and their similarity agrees with a previous comparison (Stöck et al., 2021). This suggests an investigation of the range of nuclear divergence for various parental species pairs.
Data availability
Custom scripts used and the data underlying this article including RAD-seq reads, SNP datasets, BLAST query sequences, and sequence datasets of nuclear and mitochondrial loci are available on ScienceDB (https://doi.org/10.57760/sciencedb.14822) (will be published upon final acceptance of the manuscript; data private access link: https://www.scidb.cn/en/s/URnIba). Mitochondrial sequences obtained by the Sanger method were deposited in GenBank under accessions PP065924–36 and PP083013–178.
Supplementary Files

ADMIXTURE cross-validation plot and ancestry proportions for 55 Quasipaa boulengeri individuals based on the 45,523-SNP dataset.
*: Sample 28B.


























Statistical parsimony networks for haplotypes of 26 loci.
Corresponding scaffolds of published genomes of Quasipaa spinosa (Qs) and Nanorana parkeri (Np) are shown. The 0a and 1a in the sequence name indicate allele 0 and allele 1, respectively. The last character in a sequence name, b, r, or H, indicates Q. boulengeri, Q. robertingeri, or putative hybridogenetic individuals, respectively. Between them, the sample ID is given. Lines represent single substitutions, and numbers beside lines are site numbers. Circles are haplotypes not encountered.

Neighbor-joining tree based on the SNP identity-by-state distance matrix for individuals.
Numbers at nodes are bootstrap support values ≥70, and circles indicate support values of 100. BB: Qu-asipaa boulengeri; RR: Q. robertingeri; BR: putative hybridogenetic hybrids between Q. boulengeri and Q. robertingeri.
Acknowledgements
We thank Pizhu Zhang and Siqi Yuan for their help with this work and thank Jinzhong Fu for discussions. We also thank three anonymous reviewers for their helpful comments on a previous version of the manuscript. We thank editors for their comments helping to improve the manuscript. This work was conducted with the permission of the Ethics Committee of the Chengdu Institute of Biology (CIBDWLL2021009) and was supported by the National Natural Sciences Foundation of China (NSFC-32170465 to Y.Z., NSFC-32170419 to X.Z., and NSFC-32270448 to Y.X.).
Additional files
References
- Fast model-based estimation of ancestry in unrelated individualsGenome Research 19:1655–1664Google Scholar
- FastQC: a quality control tool for high throughput sequence datahttp://www.bioinformatics.babraham.ac.uk/projects/fastqc/
- Meiosis and gametogenesis in hybrid, polyploid, and clonal fishes: case studies in the dojo loach Misgurnus anguillicaudatusFisheries Science 89:537–570Google Scholar
- Clonality: the genetics, ecology, and evolution of sexual abstinence in vertebrate animalsNew York: Oxford University Press Google Scholar
- The evolutionary network of whiptail lizards reveals predictable outcomes of hybridizationScience 377:773–777Google Scholar
- Stable species boundaries despite ten million years of hybridization in tropical eelsNature Communications 11:1433Google Scholar
- Embryonal and larval development of F1 generation of green frogs’ different combinationsActa Zoologica Cracoviensia 12:123–160Google Scholar
- Species boundaries in carp gudgeons (Eleotrididae: Hypseleotris) from the River Murray, South Australia: evidence for multiple species and extensive hybridizationMarine and Freshwater Research 51:805–815Google Scholar
- Time and time again: unisexual salamanders (genus Ambystoma) are the oldest unisexual vertebratesBMC Evolutionary Biology 10:238Google Scholar
- 110 years of orthopteran cytogenetics, the chromosomal evolutionary viewpoint, and Michael White’s signal contributions to the fieldJournal of Orthoptera Research 19:165–182Google Scholar
- Sex in unisexual salamanders: discovery of a new sperm donor with ancient affinitiesHeredity 103:483–493Google Scholar
- Unisexual salamanders (genus Ambystoma) present a new reproductive mode for eukaryotesGenome 50:119–136Google Scholar
- Trimmomatic: a flexible trimmer for Illumina sequence dataBioinformatics 30:2114–2120Google Scholar
- Heterosis and interclonal variation in thermal tolerance in unisexual fishesEvolution 33:848–859Google Scholar
- Speciation by hybridization in phasmids and other insectsCanadian Journal of Zoology 68:1747–1760Google Scholar
- BBMap: a fast, accurate, splice-aware alignerhttps://sourceforge.net/projects/bbmap/
- BLAST+: architecture and applicationsBMC Bioinformatics 10:421Google Scholar
- Clustering mitochondrial DNA sequences experienced tandem duplication based on alignment-free comparison in Quasipaa boulengeriSichuan Journal of Zoology 37:261–267Google Scholar
- Hybridogenetic reproduction and maternal ancestry of polyploid Iberian fish: the Tropidophoxinellus alburnoides complexGenetics 146:983–993Google Scholar
- Stacks: an analysis tool set for population genomicsMolecular Ecology 22:3124–3140Google Scholar
- Second generation PLINK: rising to the challenge of larger and richer datasetsGigaScience 4:7Google Scholar
- Phylogeny of the Asian spiny frog tribe Paini (Family Dicroglossidae) sensu DuboisMolecular Phylogenetics and Evolution 50:59–73Google Scholar
- TBtools: an integrative toolkit developed for interactive analyses of big biological dataMolecular Plant 13:1194–1202Google Scholar
- TCS: a computer program to estimate gene genealogiesMolecular Ecology 9:1657–1660Google Scholar
- Reconfirming the hybrid origin and generic status of the Iberian cyprinid complex Squalius alburnoidesJournal of Fish Biology 76:707–715Google Scholar
- The variant call format and VCFtoolsBioinformatics 27:2156–2158Google Scholar
- Mutual maintenance of di-and triploid Pelophylax esculentus hybrids in R-E systems: results from artificial crossings experimentsBMC Evolutionary Biology 17:220Google Scholar
- Capture and return of sexual genomes by hybridogenetic frogs provides clonal genome enrichment in a sexual speciesScientific Reports 11:1633Google Scholar
- An extinct vertebrate preserved by its living hybridogenetic descendantScientific Reports 7:12768Google Scholar
- Population genomics of an exceptional hybridogenetic system of Pelophylax water frogsBMC Evolutionary Biology 19:164Google Scholar
- Hybridogenesis in water frogseLS 1:718–726Google Scholar
- Multiple uprising invasions of Pelophylax water frogs, potentially inducing a new hybridogenetic complexScientific Reports 7:6506Google Scholar
- Piecing the barcoding puzzle of Palearctic water frogs (Pelophylax) sheds light on amphibian biogeography and global invasionsGlobal Change Biology 30:e17180Google Scholar
- ipyrad: Interactive assembly and analysis of RADseq datasetsBioinformatics 36:2592–2594Google Scholar
- Atlas of amphibians in China, field ed. ZhengzhouHenan Science and Technology Press Google Scholar
- Fauna Sinica. Amphibia. Volume 3. Anura RanidaeBeijing: Science Press Google Scholar
- PHYLIP (phylogeny inference package) version 3.695Department of Genome Sciences and Department of Biology, University of Washington http://evolution.genetics.washington.edu/phylip/
- Spontaneous mutation accumulation in Daphnia pulex in selection-free vs. competitive environmentsMolecular Biology and Evolution 34:160–173Google Scholar
- Amphibian species of the world: an online reference. Version 6.2. American Museum of Natural Historyhttps://amphibiansoftheworld.amnh.org/
- The highest-elevation frog provides insights into mechanisms and evolution of defenses against high UV radiationProceedings of the National Academy of Sciences of the United States of America 119:e2212406119Google Scholar
- Evolutionary dynamics and consequences of parthenogenesis in vertebratesAnnual Review of Ecology, Evolution, and Systematics 51:191–214Google Scholar
- Sympatric hybridization and allozyme variation in the toads Bufo americanus and B. fowleri in southern OntarioCopeia 1984:18–26Google Scholar
- Reconstructing mitochondrial genomes directly from genomic next-generation sequencing reads—a baiting and iterative mapping approachNucleic Acids Research 41:e129Google Scholar
- A Robertsonian translocation suppresses a somatic recombination pathway to loss of heterozygosityNature Genetics 33:33–39Google Scholar
- Reproductive multitasking: the female gametophyteAnnual Review of Plant Biology 71:517–546Google Scholar
- Mitochondrial genome organization and divergence in hybridizing central European waterfrogs of the Pelophylax esculentus complex (Anura, Ranidae)Gene 491:71–80Google Scholar
- Spontaneous heterosis in larval life-history traits of hemiclonal frog hybridsProceedings of the National Academy of Sciences of the United States of America 96:2171–2176Google Scholar
- Quasipaa spinosa’s genome assembly and annotation filesDryad https://doi.org/10.5061/dryad.ghx3ffbpw
- A chromosomal level genome sequence for Quasipaa spinosa (Dicroglossidae) reveals chromosomal evolution and population diversityMolecular Ecology Resources 22:1545–1558Google Scholar
- Defining loci in restriction-based reduced representation genomic data from nonmodel species: sources of bias and diagnostics for optimal clusteringBioMed research international 2014:e675158Google Scholar
- The IUCN Red List of Threatened Specieshttps://www.iucnredlist.org
- Divergent sperm factors for egg activation in amphibian fertilizationReproduction 164:F29–F37Google Scholar
- Genomic features of parthenogenetic animalsJournal of Heredity 112:19–33Google Scholar
- Complete mitochondrial genome of the headwater livebearer, Poeciliopsis monacha: the mother of clonesMitochondrial DNA Part B 1:793–794Google Scholar
- Maternal identification of hybrid eggs in Hexagrammos spp. by means of multiplex amplified product length polymorphism of mitochondrial DNAAquatic Biology 1:187–194Google Scholar
- Identification of hemiclonal reproduction in three species of Hexagrammos marine reef fishesJournal of Fish Biology 85:189–209Google Scholar
- The RadOrgMiner pipeline: automated genotyping of organellar loci from RADseq dataMethods in Ecology and Evolution 13:1962–1975Google Scholar
- Maternal ploidy shapes reproductive pathways in the triploid hybrid Chrosomus eos x eos-neogaeusMolecular Ecology 33:e17264Google Scholar
- Unexpected oogenic pathways for the triploid fish Chrosomus eos-neogaeusJournal of Heredity 110:370–377Google Scholar
- A little bit is better than nothing: the incomplete parthenogenesis of salamanders, frogs and fishBMC Biology 8:78Google Scholar
- PartitionFinder 2: new methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analysesMolecular Biology and Evolution 34:772–773Google Scholar
- Clustal W and Clustal X version 2.0Bioinformatics 23:2947–2948Google Scholar
- Naturally clonal vertebrates are an untapped resource in ecology and evolution researchNature Ecology & Evolution 3:161–169Google Scholar
- HybridogenesisCurrent Biology 29:R9–R11Google Scholar
- Evolution of alternative reproductive systems in Bacillus stick insectsEvolution 78:1109–1120Google Scholar
- Evolutionary and ecological implications of sexual parasitismTrends in Ecology & Evolution 28:297–306Google Scholar
- Fast and accurate short read alignment with Burrows-Wheeler transformBioinformatics 25:1754–1760Google Scholar
- The Sequence Alignment/Map format and SAMtoolsBioinformatics 25:2078–2079Google Scholar
- Rampant historical mitochondrial genome introgression between two species of green pond frogs, Pelophylax nigromaculatus and P. plancyiBMC Evolutionary Biology 10:201Google Scholar
- Uniparental genome elimination in Australian carp gudgeonsGenome Biology and Evolution 13:evab030Google Scholar
- Genetic structure and phyletic relationships of eastern Mediterranean Bacillus atticus Brunner (Insecta Phasmatodea): a biochemical studyBiochemical Genetics 31:343–362Google Scholar
- The mitochondrial cytochrome oxidase II gene in Bacillus stick insects: ancestry of hybrids, androgenesis, and phylogenetic relationshipsMolecular Phylogenetics and Evolution 19:157–163Google Scholar
- Current reproductive isolation between ancestors of natural hybrids in Bacillus stick insects (Insecta: Phasmatodea)Heredity 77:261–268Google Scholar
- Restriction site-associated DNA sequencing, genotyping error estimation and de novo assembly optimization for population genetic inferenceMolecular Ecology Resources 15:28–41Google Scholar
- Independent origins of allotriploidy in the fish genus PoeciliopsisJournal of Heredity 96:32–39Google Scholar
- Draft genome assembly and annotation of the Gila topminnow Poeciliopsis occidentalisFrontiers in Ecology and Evolution 7:404Google Scholar
- Historical biogeography of the livebearing fish genus Poeciliopsis (Poeciliidae: Cyprinodontiformes)Evolution 56:972–984Google Scholar
- Functional ecology of a narrow endemic plant and a widespread congener from semiarid SpainJournal of Arid Environments 73:784–794Google Scholar
- Allele-specific expression variation at different ploidy levels in Squalius alburnoidesScientific Reports 9:3688Google Scholar
- An empirical pipeline for choosing the optimal clustering threshold in RADseq studiesMolecular Ecology Resources 19:1195–1204Google Scholar
- Endemism, species selection and the origin and distribution of the vascular plant flora of New ZealandJournal of Biogeography 28:199–216Google Scholar
- The effects of translocations on recombination frequency in Caenorhabditis elegansGenetics 120:987–1001Google Scholar
- When can chromosomes drive speciation? The peculiar case of the Malagasy tomato frogs (genus Dyscophus)Zoologischer Anzeiger 268:41–46Google Scholar
- Lizards as model organisms of sex chromosome evolution: What we really know from a systematic distribution of available data?Genes 12:1341Google Scholar
- All-female strains of the teleost fishes of the genus PoeciliopsisScience 130:1656–1657Google Scholar
- The distribution of individual heterozygosity in natural populationsGenetics 95:1643–1654Google Scholar
- First empirical evidence of naturally occurring androgenesis in vertebratesRoyal Society Open Science 4:170200Google Scholar
- Origins of two hemiclonal hybrids among three Hexagrammos species (Teleostei: Hexagrammidae): genetic diversification through host switchingEcology and Evolution 6:7126–7140Google Scholar
- The protected tree Dimorphandra wilsonii (Fabaceae) is a population of inter-specific hybrids: recommendations for conservation in the Brazilian Cerrado/Atlantic Forest ecotoneAnnals of Botany 126:191–203Google Scholar
- Genetic structure of bisexual and parthenogenetic populations of Artemia from Italian brackish–hypersaline watersOceanologica Acta 26:93–100Google Scholar
- Estimation of average heterozygosity and genetic distance from a small number of individualsGenetics 89:583–590Google Scholar
- GenAlEx 6.5: genetic analysis in Excel. Population genetic software for teaching and research—an updateBioinformatics 28:2537–2539Google Scholar
- Inference of population splits and mixtures from genome-wide allele frequency dataPLoS Genetics 8:e1002967Google Scholar
- Widespread unidirectional transfer of mitochondrial DNA: a case in western Palaearctic water frogsJournal of Evolutionary Biology 21:668–681Google Scholar
- Biogeographic analysis reveals ancient continental vicariance and recent oceanic dispersal in amphibiansSystematic Biology 63:779–797Google Scholar
- A de novo case of floating chromosomal polymorphisms by translocation in Quasipaa boulengeri (Anura, Dicroglossidae)PLoS ONE 7:e46163Google Scholar
- BEDTools: a flexible suite of utilities for comparing genomic featuresBioinformatics 26:841–842Google Scholar
- Stacks 2: analytical methods for paired-end sequencing improve RADseq-based population genomicsMolecular Ecology 28:4737–4754Google Scholar
- Linkage between sexual and asexual lineages: genome evolution in Bacillus stick insectsBiological Journal of the Linnean Society 79:137–150Google Scholar
- Hybridogenesis and a potential case of R2 non-LTR retrotransposon horizontal transmission in Bacillus stick insects (Insecta Phasmida)Scientific Reports 7:41946Google Scholar
- The evolutionary ecology of gynogenesisAnnual Review of Ecology, Evolution, and Systematics 36:399–417Google Scholar
- Cytonuclear evidence for hybridogenetic reproduction in natural populations of the Australian carp gudgeon (Hypseleotris: Eleotridae)Molecular Ecology 20:3367–3380Google Scholar
- Hybridization, unisexuality and polyploidy in the teleost Poeciliopsis (Poeciliidae) and other vertebratesThe American Naturalist 103:605–619Google Scholar
- Androgenesis: where males hijack eggs to clone themselvesPhilosophical Transactions of the Royal Society B: Biological Sciences 371:20150534Google Scholar
- The complete mitochondrial genome of Quasipaa boulengeri (Anura: Dicroglossidae)Mitochondrial DNA 25:83–84Google Scholar
- SeqKit: a cross-platform and ultrafast toolkit for FASTA/Q file manipulationPLoS ONE 11:e0163962Google Scholar
- RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogeniesBioinformatics 30:1312–1313Google Scholar
- Sex chromosomes in meiotic, hemiclonal, clonal and polyploid hybrid vertebrates: along the ‘extended speciation continuum’Philosophical Transactions of the Royal Society B 376:20200103Google Scholar
- SLAF-seq: an efficient method of large-scale de novo SNP discovery and genotyping using high-throughput sequencingPloS One 8:e58700Google Scholar
- Karyological evidence of hybridogenesis in Greenlings (Teleostei: Hexagrammidae)PLoS ONE 12:e0180626Google Scholar
- Unisexual hybrids break through an evolutionary dead end by two-way backcrossingEvolution 74:392–403Google Scholar
- Polyspermic eggs in the anuran Discoglossus pictus develop normallyDevelopment 105:343–349Google Scholar
- Preliminary studies of the karyotypes of there “spine-frogs” and the karyotypic evolution of the subgenus Paa (Anura: Ranidae, Rana)Acta Herpetologica Sinica 6:35–38Google Scholar
- The birth of aposematism: high phenotypic divergence and low genetic diversity in a young clade of poison frogsMolecular Phylogenetics and Evolution 109:283–295Google Scholar
- Species delineation and systematics of a hemiclonal hybrid complex in Australian freshwaters (Gobiiformes: Gobioidei: Eleotridae: Hypseleotris)Royal Society Open Science 9:220201Google Scholar
- Comparative phylogeography of five sympatric Hypseleotris species (Teleostei: Eleotridae) in south-eastern Australia reveals a complex pattern of drainage basin exchanges with little congruence across speciesJournal of Biogeography 34:1518–1533Google Scholar
- What does the geography of parthenogenesis teach us about sex?Philosophical Transactions of the Royal Society B: Biological Sciences 371:20150538Google Scholar
- Androgenetics and triploids from an interacting parthenogenetic hybrid and its ancestors in stick insectsEvolution 50:1251–1258Google Scholar
- Genome exclusion and two strategies of chromosome duplication in oogenesis of a hybrid frogNaturwissenschaften 78:32–34Google Scholar
- Perspectives on the clonal persistence of presumed ‘ghost’ genomes in unisexual or allopolyploid taxa arising via hybridizationScientific Reports 9:4730Google Scholar
- Fixation of deleterious mutations in clonal lineages: evidence from hybridogenetic frogsEvolution 55:2319–2332Google Scholar
- Non-hybrid offspring from matings between hemiclonal hybrid waterfrogs suggest occasional recombination between clonal genomesEcology Letters 4:628–636Google Scholar
- Variation and clonal structure in a unisexual fishThe American Naturalist 112:41–55Google Scholar
- Genomic population structure of sympatric sexual and asexual populations in a parasitic wasp, Meteorus pulchricornis (Hymenoptera: Braconidae), inferred from six hundred single-nucleotide polymorphism lociMolecular Ecology 30:1612–1623Google Scholar
- Amphibian sex determination and sex reversalCellular and Molecular Life Sciences 55:901–909Google Scholar
- Chromosome research on Paa boulengeri and Paa yunnanensis (Ranidae: Paa)Chengdu: Sichuan University Google Scholar
- Mitochondrial DNA revealed the validation of Quasipaa robertingeri (Amphibia: Anura: Dicroglossidae) and its population genetic diversityMitochondrial DNA Part B 6:668–671Google Scholar
- A new ranid species of the spinosae group from SichuanIn:
- Zhao E. M.
- The complete mitochondrial genome of Quasipaa exilispinosa (Anura: Dicroglossidae)Mitochondrial DNA Part B 5:2705–2706Google Scholar
- The origin and evolution of chromosomal reciprocal translocation in Quasipaa boulengeri (Anura, Dicroglossidae)Frontiers in Genetics 10:1364Google Scholar
- Intraspecific rearrangement of mitochondrial genome suggests the prevalence of the tandem duplication-random loss (TDLR) mechanism in Quasipaa boulengeriBMC Genomics 17:965Google Scholar
- Multi-locus genetic analyses of Quasipaa from throughout its distributionMolecular Phylogenetics and Evolution 163:107218Google Scholar
- Geological events play a larger role than Pleistocene climatic fluctuations in driving the genetic structure of Quasipaa boulengeri (Anura: Dicroglossidae)Molecular Ecology 22:11201133Google Scholar
- DNA barcoding at riverscape scales: assessing biodiversity among fishes of the genus Cottus (Teleostei) in northern Rocky Mountain streamsMolecular Ecology Resources 13:583–595Google Scholar
- Suppressed recombination of sex chromosomes is not caused by chromosomal reciprocal translocation in spiny frog (Quasipaa boulengeri)Frontiers in Genetics 9:288Google Scholar
- Complete mitochondrial genomes of Nanorana taihangnica and N. yunnanensis (Anura: Dicroglossidae) with novel gene arrangements and phylogenetic relationship of DicroglossidaeBMC Evolutionary Biology 18:26Google Scholar
- UCE Phylogenomics, detection of a putative hybrid population, and one older mitogenomic node age of Batrachuperus salamandersMolecular Phylogenetics and Evolution 163:107239Google Scholar
- Dynamics behind disjunct distribution, hotspot-edge refugia, and discordant RADseq/mtDNA variability: insights from the Emei mustache toadBMC Evolutionary Biology 20:111Google Scholar
Article and author information
Author information
Version history
- Preprint posted:
- Sent for peer review:
- Reviewed Preprint version 1:
Cite all versions
You can cite all versions using the DOI https://doi.org/10.7554/eLife.109069. This DOI represents all versions, and will always resolve to the latest one.
Copyright
© 2025, Zheng et al.
This article is distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use and redistribution provided that the original author and source are credited.
Metrics
- views
- 142
- downloads
- 0
- citations
- 0
Views, downloads and citations are aggregated across all versions of this paper published by eLife.