Skip to main content
PLOS One logoLink to PLOS One
. 2022 Oct 10;17(10):e0266430. doi: 10.1371/journal.pone.0266430

Genomic insights into the evolutionary relationships and demographic history of kiwi

Michael V Westbury 1,*, Binia De Cahsan 1, Lara D Shepherd 2, Richard N Holdaway 3,4,5, David A Duchene 1, Eline D Lorenzen 1,*
Editor: Susanne P Pfeifer6
PMCID: PMC9550048  PMID: 36215252

Abstract

Kiwi are a unique and emblematic group of birds endemic to New Zealand. Deep-time evolutionary relationships among the five extant kiwi species have been difficult to resolve, in part due to the absence of pre-Quaternary fossils to inform speciation events. Here, we utilise single representative nuclear genomes of all five extant kiwi species (great spotted kiwi, little spotted kiwi, Okarito brown kiwi, North Island brown kiwi, and southern brown kiwi) and investigate their evolutionary histories with phylogenomic, genetic diversity, and deep-time (past million years) demographic analyses. We uncover relatively low levels of gene-tree phylogenetic discordance across the genomes, suggesting clear distinction between species. However, we also find indications of post-divergence gene flow, concordant with recent reports of interspecific hybrids. The four species for which unbiased levels of genetic diversity could be calculated, due to the availability of reference assemblies (all species except the southern brown kiwi), show relatively low levels of genetic diversity, which we suggest reflects a combination of older environmental as well as more recent anthropogenic influence. In addition, we suggest hypotheses regarding the impact of known past environmental events, such as volcanic eruptions and glacial periods, on the similarities and differences observed in the demographic histories of the five kiwi species over the past million years.

Introduction

New Zealand’s unique and emblematic kiwi, comprising five extant species, represent a highly divergent avian lineage within Palaeognathae. Kiwi are the only members of the Apterygidae family, and are endemic to New Zealand. The species display a number of unusual biological attributes more commonly associated with small mammals. These include low metabolic rate, lack of colour vision, flightlessness, increased longevity, and nocturnality [1, 2].

The evolutionary origins of kiwi are yet to be fully elucidated. Kiwi are estimated to have diverged from their nearest known relative, the elephant birds (Aepyornithidae), ~50 million years ago (Ma) [3]. Elephant birds are only known from Madagascar, and it is therefore thought kiwi arrived in New Zealand via flight rather than Gondwanan continental movement and vicariance [4]. Despite such a deep divergence from its closest known relative, the oldest reported kiwi fossil has been dated to at least 19–16 Ma [5], and molecular data have suggested the root of living kiwi can be traced back to between ~15 Ma [6] and ~5 Ma [7]. Thus, wide uncertainty remains with regard to the lineage leading to extant kiwi, after Apterygidae diverged from Aepyornithidae. Furthermore, deep evolutionary relationships are obscured by the lack of pre-Quaternary fossils to inform speciation events among the extant species.

The five recognized species of kiwi all belong to the genus Apteryx. The species are placed into two morphologically and genetically distinct clades [7, 8] (Fig 1). One clade comprises two extant species, great spotted kiwi (A. maxima, formerly A.haastii [9]) and little spotted kiwi (A. owenii). The other clade includes Okarito brown kiwi (A. rowi, also known as rowi), southern brown kiwi (A. australis, also known as tokoeka), and North Island brown kiwi (A. mantelli). However, despite the deep divergences of these clades and lineages [3, 6, 7], interspecific hybridization between species from different clades has recently been reported [9].

Fig 1. Phylogenetic relationships and distribution ranges of the five extant kiwi species.

Fig 1

Modern and historic distributions are based on [51]. Lightly shaded areas show estimated distributions prior to human arrival and dark shading shows current distribution. Current distribution for the little spotted kiwi is not shown, as it is presently only found on small offshore islands and wildlife sanctuaries. Population size estimates are shown in parentheses under each species name [51].

Prior to the arrival of humans 1,000–800 years ago, kiwi were found from coastal to subalpine habitats across New Zealand, but were most common in lowland rainforest habitats [7]. Partly due to their unique traits, such as flightlessness and longevity, kiwi are highly vulnerable to mammalian predators. Since human colonisation, the distribution ranges of all five extant kiwi species have been greatly reduced (Fig 1). The conservation status of four of the five species is relatively critical (S1 Table in S1 File), and is attributed to population declines following human colonisation [10, 11].

Previous studies using short fragments of mitochondrial DNA retrieved from prehistoric and contemporary specimens revealed the genetic consequences of these past anthropogenically driven range contractions, and suggested significant loss of genetic diversity after the arrival of humans, especially in the Okarito brown, southern brown, and little spotted kiwi [11, 12]. In fact, all extant little spotted kiwi are believed to descend from five founding individuals that were placed on the predator-free Kapiti Island in 1912 [13].

The use of genome-wide data can greatly enhance our existing understanding of kiwi relationships, as they allow investigation of species’ deep-time evolutionary histories. De novo nuclear genome assemblies are available from four of the five kiwi species (great spotted kiwi, little spotted kiwi, Okarito brown kiwi, North Island brown kiwi) [2, 14], and resequencing data are available from the southern brown kiwi [15]. Recent studies incorporating genomic data from kiwi have begun to address outstanding questions regarding demographic histories and divergences within kiwi. Palaeognaths have been reported to show large phylogenetic discordance across the genome, although the study did not focus on kiwi specifically [14]. Studies focussed on kiwi have reported monophyletic relationships among species, albeit with some signals of admixture between individual lineages [7, 15]. Furthermore, based on demographic history reconstructions of the past 100 thousand years it has been suggested changes in kiwi effective population sizes were likely driven by past changes in glacial extent [7, 15].

While these studies provide valuable insights into the evolutionary history of kiwi, an investigation of the presence of gene flow between lineages, and insights into deeper time (up to 1 Ma) demographic trends, is lacking. Here, we utilise single genome representatives of each extant kiwi species to better understand the interspecific evolutionary relationships among species, and to elucidate how past environmental changes may have impacted their demographic histories. Furthermore, due to the availability of reference genomes for four out of the five kiwi species, we evaluate the resilience of our phylogenomic and demographic results to reference genome biases.

Materials and methods

Data

We downloaded raw Illumina sequencing reads and genome assemblies from each of the four available kiwi species from Genbank: great spotted kiwi, little spotted kiwi, Okarito brown kiwi, and North Island brown kiwi [2, 14]. We downloaded the raw reads for a southern brown kiwi from the individual with the highest number of available reads [15]. To act as an outgroup for the phylogenomic analyses, we downloaded raw sequencing reads and the genome assembly of emu (Dromaius novaehollandiae) [14]. For estimates of comparative genetic diversity, we downloaded raw sequencing reads and the genome assemblies of three additional Paleognath species: ostrich (Struthio camelus) [16], southern cassowary (Casuarius casuarius) [14], and greater rhea (Rhea americana) [14]. Accession code details for all species can be found in S2 Table in S1 File and assembly statistics in S3 Table in S1 File.

Data processing

We trimmed adapter sequences and removed reads shorter than 30 bp from the downloaded raw reads using skewer [17], mapped the trimmed reads to the specified reference genome assembly using BWA v0.7.15 [18] and the mem algorithm. We parsed the output and removed duplicates with SAMtools v1.6 [19]. Furthermore, as some downstream analyses required the removal of sex chromosomes, we independently determined which scaffolds were most likely autosomal in origin for each reference genome assembly used in the current study. We found putative sex chromosome scaffolds by aligning each reference genome assembly to the chicken (Gallus gallus) Z and W chromosomes (Genbank accessions: Z—CM000122.5, W—CM000121.5). Alignments were performed using satsuma synteny [20] and default parameters. Information on all mappings performed can be found in S4 Table in S1 File.

Phylogenomics

We performed multiple phylogenomic analyses using different mapping reference genome assemblies, to assess the robustness of our results to reference choice. We independently mapped the raw reads of all five kiwi species to two reference genome assemblies: emu and great spotted kiwi. We also mapped the emu raw reads to the emu reference genome assembly. From the resultant mapped files, we built majority rules consensus sequences (-doFasta 2) in ANGSD v0.921 with the following parameters; -mininddepth 5 -minmapq 30 -minq 30 -uniqueonly 1, and only included autosomal scaffolds >100kb in length (-rf). We extracted 2 kb windows with a 1 Mb slide from each consensus sequence using bedtools v2.26.0 [21]. This gave us 1,546 windows when mapping to the emu and 1,889 windows when mapping to the great spotted kiwi. We filtered for windows that had at least 10 parsimony-informative sites and performed maximum likelihood phylogenetic inference for each remaining window in IQ-TREE2 [22], using the best GTR+R+F model according to the Bayesian information criterion [23]. Using the estimated gene trees (which in this case refers to a window), we inferred the species tree of kiwi under each reference genome assembly, using the summary multispecies coalescence as implemented in ASTRAL v4.10 [24]. The species tree inferred for each of the two datasets was used as the focal tree to calculate gene- and site-concordance factors for each branch in IQ-TREE2 [25].

Quantifying introgression via branch lengths

To investigate whether phylogenetic discordance among all possible kiwi triplets [[A,B],C] can be explained by incomplete lineage sorting (ILS) alone, or by a combination of ILS and gene flow, we implemented Quantifying Introgression via Branch Lengths (QuIBL) [26] on the dataset obtained when mapping the kiwi and emu raw read data to the emu reference genome. We used the kiwi and emu mapped to the emu dataset due to the requirement of an outgroup. Furthermore, since this analysis relies on the relative ages of nodes, we only used loci with nearly constant evolutionary rates among lineages. Specifically, we rooted each gene tree with the emu and excluded those loci with a coefficient of variation in root-to-tip length >0.01. We ran QuIBL specifying the emu as the overall outgroup (totaloutgroup), to test either ILS or ILS with gene flow (numdistributions 2), the number of total EM steps as 50 (numsteps), and a likelihood threshold of 0.01. We determined the significance of gene flow by comparing values of BIC1 (ILS alone) and BIC2 (ILS and gene flow). If the difference between BIC1 and BIC2 was greater than 10, as suggested by the original paper describing the method [26], we assumed incongruent topologies arose due to both ILS and gene flow. With a difference of less than 10, we assumed ILS alone.

f-branch statistic

To further investigate the potential for gene flow between kiwi lineages, we implemented the f-branch test [27, 28] The test takes correlated allele sharing into account when visualising ƒ4-ratio results meaning it can also uncover indications of gene flow between ancestral lineages. As input we created a multi-individual variant call file (VCF) using the data mapped to the emu genome and BCFtools v1.6 [29]. Specifically we used BCFtools mpileup, and filtered the VCF file to only include autosomal SNPs using BCFtools call and the -mv parameter. We ran the multi-individual VCF through Dtrios in Dsuite v0.4 r43 [27] and specified the species tree (Fig 1) and otherwise default parameters. We ran the output from Dtrios through f-branch and visualised the output using the dtools.py script from Dsuite. The default parameters for the f-branch statistic in Dsuite only consider fb, a value indicating excess allele sharing between a given branch (relative to its sister branch) and a non-sister branch, with p<0.01. However, we also assessed statistical significance of fb using a block Jack-knife approach by including the -Z parameter when running the f-branch statistic in Dsuite. We used the default number of 20 Jackknife blocks for this test. A Z score |Z|>3 was considered as significant.

End of lineage sorting/gene flow

To estimate the point in time when the genomes of the five kiwi species had fully coalesced, putatively indicating an end of lineage sorting and/or gene flow, we used the F1 hybrid Pairwise Sequentially Markovian Coalescent model, hPSMC [30]. hPSMC utilises pseudo-diploid sequences by merging pseudo-haploid sequences from two different genomes, which in our case are from different species. As the variation in the interspecific pseudo-F1 hybrid genome cannot coalesce more recently than the emergence of reproductive isolation between the two parental species, we can use this method to infer when reproductive isolation between two species may have occurred. If some regions within the genomes of two target species are yet to fully diverge, due to ILS or to gene flow, hybridisation may still be possible.

Previous studies have shown that the results of hPSMC are not significantly influenced by mapping reference [31, 32], and we therefore performed each analysis once, with the raw reads from the five kiwi species mapped to the great spotted kiwi reference genome assembly. We constructed haploid consensus sequences for each of the kiwi individuals using the same consensus sequences as the phylogenomic analysis. We merged the resultant haploid consensus sequences pairwise into a pseudo-diploid sequence using a python script available as part of the hPSMC toolsuite. The resultant pseudo-diploid sequences were run through a Pairwise Sequentially Markovian Coalescent model (PSMC) [33]. We ran PSMC specifying standard atomic intervals (4+25*2+4+6), a maximum number of iterations of 25 (-N), maximum 2N0 coalescent time of 15 (-t), and initial theta/rho ratio of 5 (-r).

To calibrate the PSMC plots, we calculated an Apteryx average mutation rate. We did this by first calculating the average pairwise distance for each species pair (S5 Table in S1 File), and dividing that by 2 * the previously published divergence times [3]. However, it should be noted that this calculation only provides an estimate and is influenced by divergence times. Given this, we selected the divergence times from Yonezawa et al [3] because they used the dataset with the largest number of genetic markers at the time and included all five kiwi species in their estimates and we therefore deemed it the most reliable. We calculated the pairwise distances in ANGSD v0.921 [34] using a consensus base call approach, with all species mapped to the great spotted kiwi reference genome assembly, and applying the same filters as for the phylogenomic analyses with the additions of only including sites found in all individuals (-minInd 5), and print a distance matrix (-makematrix 1). Given the possibility for ILS and gene flow in the kiwi lineages, whole genome pairwise distances may be under-estimated from whole genome alignments. This resulted in a mutation rate of 8.0x10-10 per year or 2.0x10-8 per generation, assuming a generation time of 25 years [7]. Our mutation rate of 2.0e-8 per generation is comparable to the rate of 1.34e-8 per generation from Bemmel et al [15].

From the PSMC output, we manually estimated the pre-divergence Ne of each pseudodiploid genome by outputting the text file (-R) using the plot script from the PSMC toolsuite. Using the pre-divergence Ne estimated from this output, we ran simulations to infer the intervals during which the pseudodiploid genomes coalesce between each species pair using ms [35]. Simulation commands in ms were automatically produced with the hPSMC_quantify_split_time.py python script from the hPSMC toolsuite, while specifying the pre-divergence Ne and the time windows we wanted to simulate, and the remaining parameters as default. The time intervals and pre-divergence Ne for each species pair can be found in S6 Table in S1 File. We plotted the results and found the simulations with an exponential increase in Ne to be closest to the empirical data, between 1.5-fold and 10-fold the value of the pre-divergence Ne. The divergence times from these simulations were taken as the time interval during which lineage sorting and/or gene flow stopped. We considered the portion between 1.5-fold and 10-fold the value of the pre-divergence Ne, as suggested in previous work [30]. This was done to capture the portion of the hPSMC plot most influenced by the divergence event. The lower bound is set to control for pre-divergence increases in population size, and the upper bound is to avoid exploring parameter space in which little information is present.

Autosome-wide heterozygosity and inbreeding estimates

We calculated autosome-wide levels of heterozygosity and runs of homozygosity (ROH) for each species from the mapped bam files using the software ROHan [36]. The raw reads of the four species for which conspecific reference genome assemblies were available, were independently mapped to each said assembly. Raw reads of the southern brown kiwi, which does not have an available conspecific assembly, were mapped to the North Island brown kiwi assembly. We ran ROHan three times independently, using different window sizes (500 kb, 1 Mb, 2 Mb), while keeping the other parameters as default. The default parameters specify a window as being a ROH if it has an average heterozygosity of less than 1x10-5. We calculated autosome-wide heterozygosity for the non-kiwi palaeognath species (emu, ostrich, southern cassowary, greater rhea) in the same way, but using only a 1 Mb window size.

We estimated the number of generations since inbreeding occurred that each respective ROH length represents using g = 100/(2rL) [37], where r = recombination rate, L = length of ROH in Mb, and g = number of generations. As genome-wide recombination rates for kiwi are unavailable, we present results based on two, 3cM per Mb based on the chicken [38] and 2.1 cM/Mb based on the rhea [39] Given this calculation, ROH>0.5Mb equates to inbreeding occurring within the last 47.6 or 33.3 generations, ROH>1Mb equates to inbreeding occurring within the last 23.8 or 16.6 generations, and ROH>2Mb equates to inbreeding occurring within the last 11.9 or 8.3 generations.

Demographic reconstruction

We ran demographic analysis on diploid consensus genomes from each kiwi species, each mapped to their conspecific reference genome assembly and the southern brown kiwi mapped to the North Island brown kiwi, using PSMC [33]. We called diploid genome sequences using SAMtools and BCFtools, specifying a minimum quality score of 20 and minimum coverage of 10. However, we also created a diploid consensus genome with Okarito brown kiwi mapped to the North Island brown kiwi, as it has previously been shown that mapping to the Okarito brown kiwi reference genome assembly may be problematic for PSMC analysis [40]. As Prasad et al. [40] did not assess the reliability of the little spotted kiwi as mapping reference, we further assessed the reliability of the little spotted kiwi PSMC results by mapping to the conspecific reference genome assembly, as well as to the great spotted kiwi and North Island brown kiwi reference genome assemblies. Before running PSMC, we removed scaffolds found to align to sex chromosomes in the previous step, and removed scaffolds shorter than 100 kb. We ran PSMC specifying the same parameters as in the hPSMC analysis and performed 100 bootstrap replicates to investigate support for the resultant demography. PSMC results were checked for overfitting so that after 20 rounds of iterations, at least 10 recombinations are inferred to have occurred in the intervals each parameter spans. We plotted the output using a mutation rate of 2.0x10-8 per generation, assuming a generation time of 25 years as detailed above for the hPSMC analysis.

Results

Mapping results

The number of reads mapping as well as the average genome-wide coverages slightly changed depending on whether individuals were mapped to a con- or heterospecific reference genome, with lower values being recovered when mapping to a heterospecific reference (S4 Table in S1 File). When mapping to conspecific reference genomes, we recovered genome-wide coverages of 24.52x for the great spotted kiwi, 20.8x for the North Island brown kiwi, 31.6x for the little spotted kiwi, and 25.2x for the Okarito brown kiwi. Mapping the southern brown kiwi to the North Island brown kiwi reference genome gave 14.8x.

Interspecific evolutionary relationships

Reads more similar to the reference map more successfully than divergent reads, artificially inflating signals of genetic similarities between a highly divergent outgroup and an ingroup species used as mapping reference [41]. As our outgroup (emu) is highly divergent from the ingroup kiwi species, our results when mapping to the ingroup great spotted kiwi could therefore be biased by the emu artificially appearing more genetically similar to the great spotted kiwi than the non-reference species. Although fewer reads mapped successfully (S4 Table in S1 File), by also mapping to the outgroup we tested for the influence of this potential bias. Our results are unlikely to be driven by reference bias as, regardless of the genome assembly selected as mapping reference, the vast majority of windows in our phylogenomic analysis support the grouping of little spotted kiwi and great spotted kiwi as sister species, and the grouping of Okarito brown kiwi, North Island brown kiwi, and southern brown kiwi in a separate clade, with the southern brown kiwi being sister to the Okarito brown kiwi, North Island brown clade (Fig 1, S7 Table in S1 File). Despite the overall support for the species tree, within brown kiwi, alternative gene tree topologies occurred at relatively high frequency compared to other alternative topologies found (up to 34.9%). Site concordance factors also found high levels of discordance within this clade (up to 42.1%).

Further examination using QuIBL to assess if alternative topologies may have arisen due to ILS alone, or to both ILS and gene flow, indicated a mixture of both ILS and gene flow causing the topological discordances in our dataset (S8 Table in S1 File). QuIBL suggests gene flow between great spotted kiwi and the Okarito brown kiwi, great spotted kiwi and southern brown kiwi, North Island brown kiwi and southern brown kiwi, Okarito brown kiwi and little spotted kiwi, and southern brown kiwi and little spotted kiwi.

The f-branch statistic suggests several signals of gene flow between kiwi lineages, as shown by elevated fb. We find elevated fb between i) the little spotted kiwi and all three brown kiwi species, ii) the southern brown kiwi and both the great and lesser spotted kiwi, and iii) the Okarito brown kiwi with both spotted kiwi and the southern brown kiwi (Fig 2). All results were significant with Z>3 (S9 Table in S1 File).

Fig 2. Genome-wide f-branch results.

Fig 2

(A) Species tree; (B) and (C) Species tree in expanded form, with internal branches as dotted lines. The values in the matrix refer to excess allele sharing between the expanded tree branch (relative to its sister branch) and the species on the x-axis. Grey squares are comparisons that could not be made due to the topological input requirements of the test and its inability to infer gene flow between sister lineages. Lines connecting branches show: (B) gene flow events inferred directly from the f-branch results. Dark purple coloured lines show results supported by QuIBL; (C) gene flow events that we hypothesise from the f-branch results, while accounting for (i) the inability to detect gene flow between sister lineages, and (ii) a lack of a positive means less gene flow relative to the sister lineage, rather than no gene flow.

To quantify when reproductive isolation may have been complete between kiwi lineages, we ran the F1 hPSMC described above. Between the two major clades, we found that lineage sorting and/or gene flow ceased 2.6–1.8 Ma. This pattern was the same, regardless of which individuals were used in the pairwise comparison (S1 Fig in S1 File). Within clades, we find complete reproductive isolation (lineage sorting was complete and/or gene flow ceased) between great spotted kiwi and little spotted kiwi 700–400 thousand years ago (kya) (S2 Fig in S1 File), and similarly between North Island brown kiwi and Okarito brown kiwi 800–500 kya (S3 Fig in S1 File). Lineage sorting was complete and/or gene flow ceased 1.2 Ma-800 kya between southern brown kiwi and both the North Island brown kiwi and the Okarito brown kiwi (S4 Fig in S1 File). hPSMC uses the PSMC method, and therefore changes in effective population size can be influenced by population structure [42] and regions undergoing strong negative selection [43]. However, in our hPSMC when comparing different species pairs, e.g. southern brown kiwi/North Island brown kiwi and southern brown kiwi/Okarito brown kiwi, we get similar if not identical results. Therefore, we do not think population structure largely influences hPSMC results. Selection on the other hand could play a role and remains to be tested. However, as previous studies showed strongly deleterious mutations mask declines in PSMC [43], and we are focusing on exponential increases, we do not think our results will be greatly influenced by this.

Autosome-wide heterozygosity and inbreeding estimates

We find the highest levels of autosome-wide heterozygosity in North Island brown kiwi, followed by great spotted kiwi, southern brown kiwi, Okarito brown kiwi, and little spotted kiwi (Fig 3A). To further contextualise these estimates, we calculated the autosome-wide heterozygosity estimates of four other palaeognaths (emu, ostrich, southern cassowary, greater rhea). Our analysis revealed that kiwi species have relatively low levels of heterozygosity compared to the other palaeognaths, except for the southern cassowary, whose heterozygosity only exceeded that of little spotted kiwi.

Fig 3. Genetic diversity and inbreeding estimates.

Fig 3

(A) Autosome-wide heterozygosity calculated using ROHan for the four kiwi species with conspecific reference genomes, and four other palaeognath species. Error bars show maximum and minimum values. (B) Autosomal runs of homozygosity calculated using various window sizes (0.5 Mb, 1 Mb, and 2 Mb) for the four kiwi species with conspecific reference genomes. Differential shading shows the window size used.

We further investigated whether recent inbreeding may have caused the low levels of autosome-wide heterozygosity using runs of homozygosity (ROH) analyses. As expected, the number of ROH increased when specifying increasingly smaller window sizes as a ROH (Fig 3B). Moreover, the general trend reflected what would be expected based on autosome-wide heterozygosity levels. That is, the species with the lowest overall heterozygosity (little spotted kiwi) also had the most ROH, and the species with the highest overall heterozygosity (North Island brown kiwi) had the least ROH. We found the lowest levels of ROH in the southern brown kiwi (Fig 3B).

This pattern was not as obvious in the other palaeognath species. The individual with the highest levels of diversity (rhea) also showed considerable levels of ROH (1 Mb window size; 2.2% of the genome in ROH). However, the individual with the lowest diversity (southern cassowary), also had the highest levels of ROH (1 Mb window size; 8.42% of the genome in ROH). The remaining two species had low levels of ROH: emu—0.50% and ostrich—0.64%.

Intraspecific demographic histories

Our assessment of how mapping reference influences PSMC results of little spotted kiwi showed similar results to those reported previously for Okarito brown kiwi [40]; we find relatively consistent results when mapping the little spotted kiwi raw reads to both North Island brown kiwi and to great spotted kiwi, but much reduced effective population size (Ne) values when mapping to little spotted kiwi (S5 Fig in S1 File). Therefore, we based our inferences on the results retrieved when mapping to its closest relative (great spotted kiwi).

We find Ne in North Island brown kiwi was relatively high and stable over the past 1 Ma, with several fluctuations within the past 100–50 kya (Fig 4). However, the bootstrap support values during the latter time period show considerable uncertainty, suggesting little confidence in these fluctuations. The southern brown kiwi also exhibited relatively stable Ne values until ~200 kya where there was a continual decrease in Ne until present. Ne in Okarito brown kiwi decreased 1–0.5 Ma, followed by a plateau and slight increase, then a more rapid decline in the past ~100 kya. The two remaining species, little spotted kiwi and great spotted kiwi, exhibit similar demographic trends. Both species show a continuous gradual decline in Ne over the past 1 Ma.

Fig 4. Effective population sizes over the past one million years calculated using PSMC.

Fig 4

Faded lines show the 100 bootstrap replicates produced for each species.

Discussion

Phylogenetic relationships

Based on our phylogenetic analysis, we found relationships among the five kiwi species to be generally conserved across windows, consistent with previous findings based on mtDNA sequences and nuclear data [3, 7, 15, 44] (Fig 1, S7 Table in S1 File). However, despite the clear overall species tree, gene (window) tree and site concordance factors both show the presence of discordant topologies across the genome, which QuIBL suggests are due to a mixture of ILS and gene flow. All signals of gene flow found in QuIBL were supported by our f-branch statistic results, to the exclusion of gene flow between the North Island and southern brown kiwis, which was not found in the f-branch statistic (Fig 2B). Furthermore, our f-branch statistic found several signals of gene flow not found with QuIBL. However, interpretations of the f-branch statistic can be complicated by results only being relative to a sister species, i.e. an fb of 0 does not mean no gene flow, but just less gene flow than the sister species. Taking this into account and the known order of divergence of each lineage [3], we hypothesise that some observed signals of gene flow may be due to retention of introgressed loci from ancestral gene flow events (Fig 2C). Little spotted kiwi showed similar fb with all three brown kiwi. This pattern could reflect independent gene flow events between little spotted kiwi and each brown kiwi species (Fig 2B), or gene flow between little spotted kiwi and the ancestral brown kiwi. Little spotted and great spotted kiwi split after southern brown split from North Island and Okarito brown kiwi, and hence little spotted kiwi alone could not have exchanged genetic material with the ancestral brown kiwi. Rather, the ancestral spotted kiwi could have exchanged genetic material with the ancestral brown kiwi, and little spotted has retained more signal of this relative to great spotted kiwi (Fig 2C). Similarly, the signal between southern brown kiwi and both the little and great spotted kiwi, as well as between the Okarito brown kiwi and both the little and great spotted kiwi, could also reflect gene flow between ancestral lineages, with the retention of less of the introgressed alleles in the North Island brown kiwi (Fig 2C). Our hPSMC results (S1-S3 Figs in S1 File) suggest coalescence between species pairs occurred relatively recently, which can be interpreted as gene flow among all species after they initially diverged. A scenario of interspecific gene flow is congruent with recent reports of hybridisation between kiwi sister species, and between kiwi species from each of the two distinct clades [9].

The presence of interspecific gene flow between kiwi species may explain the large discrepancies between molecular estimates of their divergence times. However, varying estimates could also reflect difficulties in time calibration given uncertainties in fossil dates, substitution rate, and generation time. Estimates of the stem divergence of the five extant species have differed markedly in the literature. These include, but are not limited to, ~14.5 Ma [6], ~13.4 Ma [45], ~12.31 Ma [3], ~11.3 Ma [44], ~8.5 Ma [4, 46], and ~5.9 Ma [7]. Even when considering the youngest divergence estimate (~5.9 Ma), our hPSMC results suggest the genomes completely coalesced between 2.6 Ma and 1.8 Ma, which may be interpreted as the continuation of lineage sorting for a long time post divergence, or continued gene flow.

Interspecific gene flow is well documented in avian species [47] and may be due to the relatively slow rate at which postzygotic incompatibilities accumulate in birds [48]. Therefore, premating isolation is likely an important mechanism for the maintenance of reproductive isolation in birds, although this may be hampered by high dispersal rates through flight. Kiwi are known to have evolved flightlessness independently from other palaeognaths after their arrival to New Zealand [3, 6]. However, a lack of fossils makes it difficult to reliably determine when flightlessness evolved. Nevertheless, based on a recent fossil find, kiwi are thought to be flightless from at least the Middle Pleistocene (~1 Ma) [49]. Although a lack of migration through flight likely would have limited the dispersal capabilities of kiwi, some species had partly overlapping ranges in historic times (Fig 1), which may have facilitated hybridisation.

Current and recent demography

Our finding of very low diversity in little spotted kiwi is congruent with previous studies based on microsatellite data [50] and population-level nuclear genomes [15]. Our finding—at least as represented by our sampled individual—may reflect high levels of recent inbreeding. Its high proportion of ROH >2Mb equates to inbreeding within the last ~250 years, and is congruent with current knowledge of the species. In 1912, five little spotted kiwi were moved to an offshore, predator-free island (Kapiti Island). Therefore, the entire population of the species, estimated at approximately 2,000 individuals [51] (Fig 1), is descended from at most five founders; a recent microsatellite study indicated perhaps only three founder individuals [12]. Such a low number of founders would have led to high levels of inbreeding. However, this findings contrasts with no inbreeding reported in a recent population genomic study [15]. Mapping to a phylogenetically distant reference can inflate heterozygosity estimates and remove signs of inbreeding [40]. The finding of no inbreeding on a genome-wide level [15] may be a byproduct of the computational approach employed, rather than a biological signal; the little spotted kiwi data were not mapped to a conspecific reference genome and hence the lack of inbreeding may be the result of mapping biases.

Okarito brown kiwi experienced a bottleneck of ~150 birds in the 1990s [52]. However, we find similar diversity and inbreeding levels to great spotted kiwi (Fig 3) which has an estimated population size of 14,000 [51] (Fig 1). Furthermore, based on 11 microsatellite loci, it was reported that great spotted kiwi had high levels of genetic diversity compared with other kiwi species, and showed no evidence of a recent bottleneck (within the last 100 generations) [53]. Therefore, if genetic diversity alone predicts population size, great spotted kiwi should display higher levels of diversity than Okarito brown kiwi. However, as our Okarito brown kiwi individual was sampled in 1993, our findings likely do not reflect the genetic impact of the recent bottleneck in the 1990s. Therefore, the Okarito brown kiwi population at Okarito, from which this sample arose, likely had higher levels of diversity pre-bottleneck, and exemplifies the relevance of sample age when investigating recent bottleneck events. Finally, our finding of North Island brown kiwi having the highest levels of diversity is consistent with higher census population size (25,000) than the other species [51], and is congruent with a recent population genomic study [15].

Long-term demography

Owing to a lack of fossil evidence, it is difficult to investigate how kiwi populations responded to past climatic and environmental events. Genomic data allows the exploration of changes in population size over evolutionary timescales to provide hypotheses based on correlations with known environmental events. Previous work using kiwi genomic data to investigate changes in effective population size have been interpreted within the framework of glacial and interglacial periods [7, 15].

We observe relatively stable Ne in the North Island brown kiwi over the past one million years (Fig 4). During this period, the central North Island experienced several major rhyolitic eruptions, as evidenced by tephras in ocean cores to the east and northeast of the North Island [54]. The ash from these eruptions rarely reached the northern peninsula of the North Island, providing a putative refuge for this population. Therefore, the northern distribution of North Island brown kiwi (Fig 1) may have facilitated its long-term stability in effective population size. Similarly, the southern range of the southern brown kiwi may have facilitated a stable Ne in this species 1 Ma—100 kya, despite eruptions.

In contrast, both great and little spotted kiwi show a general trend of declining Ne throughout the past one million years (Fig 4). As the genetic data of both species are from the South Island, despite the little spotted kiwi also occurring in the North Island (Fig 1), these results may only represent changes to South Island populations. Both volcanic fallout and repeated glaciations–or repeated de­glaciations–may have had long-term negative effects on both kiwi species and caused decreases in census population size and/or connectivity, which would lead to the observed decrease in Ne.

The Ne of Okarito brown kiwi also shows a decreasing trend, suggestive of similar mechanisms to those in the great and little spotted kiwi (Fig 4). However, the trajectory deviates with a plateau in Ne, before rapidly decreasing ~100 kya. Associated with this decline is the end of the last interglacial, Isotope Stage 5e (130–116 kya [55]). A change in Ne at this time could be caused by the restriction of population movement, changes in population structuring, and/or a reduction in population size, all of which could have been driven by the environmental instability of this period. However, Okarito brown kiwi has experienced many glacial and interglacial periods before this event, and therefore it is difficult to determine correlation or causation.

Although the confidence in PSMC for more recent periods of time (<20kya) is limited [33], it is notable that the decline in Ne observed across all five kiwi species within the last 40 kya broadly coincides with the most recent global volcanic super eruption (Fig 4). A decrease in Ne could be due to a decrease in absolute population size, or a decrease in connectivity between previously connected populations. Similar declines in the past 10–20 kya have previously been reported [15]. The Oruanui eruption of Taupo volcano 25.6 kya deposited 1000 km3 of tephra across New Zealand, from Auckland in the North Island to the Waitaki River in the South Island [56, 57]. Such a catastrophic event is expected to have destroyed most of the New Zealand biota between 38° and 45° south. If the decline in Ne in association with this major eruption was exhibited by just one taxon, little case could be made for cause and effect. However, the analogous temporal pattern of a population decline across all five species provides strong support for a significant environmental driver.

It is difficult to precisely pinpoint the driver of past population fluctuations, especially given the uncertainty in mutation rates and generation times. However, we form a number of hypotheses mostly based on correlations with the timing of volcanic activity. To date, the few studies using genomic data to investigate demographic histories of kiwi have alluded to glaciation as a major influence [7, 15]; although the Taupo caldera volcano ~1800 years ago has been identified as having some impact on kiwi Ne [15, 58]. By including more ancient time periods (>20kya) that experienced multiple volcanic events, we suggest volcanic eruptions may have been a previously overlooked major evolutionary force shaping the biota of New Zealand.

Supporting information

S1 File. File containing all supplementary information: S1-S9 Tables and S1-S5 Figs.

(DOCX)

Acknowledgments

We would like to acknowledge the Te Parawhau Trust and Waikaremoana iwi, Te Rūnanga o Ngāi Tahu, Te Ātiawa Manawhenua Ki Te Tau Ihu Trust, who provided guidance to the authors who generated the kiwi genomic data, on which our study is based. We would also like to thank the reviewers for their comments on improving the manuscript. The authors declare no conflicts of interest.

Data Availability

All raw data used in this manuscript were previously published and the accession codes are listed in supplementary table S2. Example commands and the scripts necessary to perform all analyses in this manuscript are available at github.com/Mvwestbury/Kiwi-genomes.

Funding Statement

This work was supported by the Independent Research Fund Denmark | Natural Sciences, Forskningsprojekt 1, grant no. 8021-00218B, and the Villum Fonden Young Investigator Programme, grant no. 13151, to EDL. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

References

  • 1.Ramstad KM, Miller HC, Kolle G. Sixteen kiwi (Apteryx spp) transcriptomes provide a wealth of genetic markers and insight into sex chromosome evolution in birds. BMC Genomics. 2016;17: 410. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Le Duc D, Renaud G, Krishnan A, Almén MS, Huynen L, Prohaska SJ, et al. Kiwi genome provides insights into evolution of a nocturnal lifestyle. Genome Biol. 2015;16: 147. doi: 10.1186/s13059-015-0711-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Yonezawa T, Segawa T, Mori H, Campos PF, Hongoh Y, Endo H, et al. Phylogenomics and Morphology of Extinct Paleognaths Reveal the Origin and Evolution of the Ratites. Curr Biol. 2017;27: 68–77. doi: 10.1016/j.cub.2016.10.029 [DOI] [PubMed] [Google Scholar]
  • 4.Mitchell KJ, Llamas B, Soubrier J, Rawlence NJ, Worthy TH, Wood J, et al. Ancient DNA reveals elephant birds and kiwi are sister taxa and clarifies ratite bird evolution. Science. 2014;344: 898–900. doi: 10.1126/science.1251981 [DOI] [PubMed] [Google Scholar]
  • 5.Worthy T, Worthy JP, Tennyson AJD, Salisbury SW, Hand SJ, Scofield RP. Miocene fossils show that kiwi (Apteryx, Apterygidae) are probably not phyletic dwarves. Paleornithological Research 2013—Proceedings of the 8th International Meeting of the Society of Avian Paleontology and Evolution. Natural History Museum Vienna; 2013. https://dspace2.flinders.edu.au/xmlui/handle/2328/35951 [Google Scholar]
  • 6.Phillips MJ, Gibb GC, Crimp EA, Penny D. Tinamous and moa flock together: mitochondrial genome sequence analysis reveals independent losses of flight among ratites. Syst Biol. 2010;59: 90–107. doi: 10.1093/sysbio/syp079 [DOI] [PubMed] [Google Scholar]
  • 7.Weir JT, Haddrath O, Robertson HA, Colbourne RM, Baker AJ. Explosive ice age diversification of kiwi. Proc Natl Acad Sci U S A. 2016;113: E5580–7. doi: 10.1073/pnas.1603795113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Shepherd LD, Worthy TH, Tennyson AJD, Scofield RP, Ramstad KM, Lambert DM. Ancient DNA analyses reveal contrasting phylogeographic patterns amongst kiwi (Apteryx spp.) and a recently extinct lineage of spotted kiwi. PLoS One. 2012;7: e42384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Shepherd LD, Tennyson AJD, Robertson HA, Colbourne RM, Ramstad KM. Hybridisation in kiwi (Apteryx; Apterygidae) requires taxonomic revision for the Great Spotted Kiwi. Avian Res. 2021;12. doi: 10.1186/s40657-021-00257-6 [DOI] [Google Scholar]
  • 10.Hurles ME, Matisoo-Smith E, Gray RD, Penny D. Untangling Oceanic settlement: the edge of the knowable. Trends Ecol Evol. 2003;18: 531–540. [Google Scholar]
  • 11.Shepherd LD, Lambert DM. Ancient DNA and conservation: lessons from the endangered kiwi of New Zealand. Mol Ecol. 2008;17: 2174–2184. doi: 10.1111/j.1365-294X.2008.03749.x [DOI] [PubMed] [Google Scholar]
  • 12.Ramstad KM, Gibb GC, Robertson HA, Colbourne RM, Doran EE, Shepherd LD. Recent extinctions among Little Spotted Kiwi (Apteryx owenii) and the origin of extant populations. Emu—Austral Ornithology. 2021; 1–10. [Google Scholar]
  • 13.Taylor H, Ramstad KM. Five kiwi that saved a species. Forest & Bird. 2012;346: 16–20. [Google Scholar]
  • 14.Sackton TB, Grayson P, Cloutier A, Hu Z, Liu JS, Wheeler NE, et al. Convergent regulatory evolution and loss of flight in paleognathous birds. Science. 2019;364: 74–78. doi: 10.1126/science.aat7244 [DOI] [PubMed] [Google Scholar]
  • 15.Bemmels JB, Mikkelsen EK, Haddrath O, Colbourne RM, Robertson HA, Weir JT. Demographic decline and lineage-specific adaptations characterize New Zealand kiwi. Proc Biol Sci. 2021;288: 20212362. doi: 10.1098/rspb.2021.2362 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Zhang G, Li C, Li Q, Li B, Larkin DM, Lee C, et al. Comparative genomics reveals insights into avian genome evolution and adaptation. Science. 2014;346: 1311–1320. doi: 10.1126/science.1251385 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Jiang H, Lei R, Ding S-W, Zhu S. Skewer: a fast and accurate adapter trimmer for next-generation sequencing paired-end reads. BMC Bioinformatics. 2014;15: 182. doi: 10.1186/1471-2105-15-182 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Li H, Durbin R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics. 2009;25: 1754–1760. doi: 10.1093/bioinformatics/btp324 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics. 2009;25: 2078–2079. doi: 10.1093/bioinformatics/btp352 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Grabherr MG, Russell P, Meyer M, Mauceli E, Alföldi J, Di Palma F, et al. Genome-wide synteny through highly sensitive sequence alignment: Satsuma. Bioinformatics. 2010;26: 1145–1151. doi: 10.1093/bioinformatics/btq102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Quinlan AR. BEDTools: The Swiss-Army Tool for Genome Feature Analysis. Curr Protoc Bioinformatics. 2014;47: 11.12.1–34. doi: 10.1002/0471250953.bi1112s47 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A, et al. IQ-TREE 2: New Models and Efficient Methods for Phylogenetic Inference in the Genomic Era. Mol Biol Evol. 2020;37: 1530–1534. doi: 10.1093/molbev/msaa015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods. 2017;14: 587–589. doi: 10.1038/nmeth.4285 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Zhang C, Rabiee M, Sayyari E, Mirarab S. ASTRAL-III: polynomial time species tree reconstruction from partially resolved gene trees. BMC Bioinformatics. 2018;19: 153. doi: 10.1186/s12859-018-2129-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Minh BQ, Hahn MW, Lanfear R. New methods to calculate concordance factors for phylogenomic datasets. Mol Biol Evol. 2020;37: 2727–2733. doi: 10.1093/molbev/msaa106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Edelman NB, Frandsen PB, Miyagi M, Clavijo B, Davey J, Dikow RB, et al. Genomic architecture and introgression shape a butterfly radiation. Science. 2019;366: 594–599. doi: 10.1126/science.aaw2090 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Malinsky M, Matschiner M, Svardal H. Dsuite—Fast D-statistics and related admixture evidence from VCF files. Mol Ecol Resour. 2021;21: 584–595. doi: 10.1111/1755-0998.13265 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Malinsky M, Svardal H, Tyers AM, Miska EA, Genner MJ, Turner GF, et al. Whole-genome sequences of Malawi cichlids reveal multiple radiations interconnected by gene flow. Nat Ecol Evol. 2018;2: 1940–1955. doi: 10.1038/s41559-018-0717-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Narasimhan V, Danecek P, Scally A, Xue Y, Tyler-Smith C, Durbin R. BCFtools/RoH: a hidden Markov model approach for detecting autozygosity from next-generation sequencing data. Bioinformatics. 2016;32: 1749–1751. doi: 10.1093/bioinformatics/btw044 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Cahill JA, Soares AER, Green RE, Shapiro B. Inferring species divergence times using pairwise sequential Markovian coalescent modelling and low-coverage genomic data. Philos Trans R Soc Lond B Biol Sci. 2016;371. doi: 10.1098/rstb.2015.0138 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Westbury MV, Petersen B, Lorenzen ED. Genomic analyses reveal an absence of contemporary introgressive admixture between fin whales and blue whales, despite known hybrids. PLoS One. 2019;14: e0222004. doi: 10.1371/journal.pone.0222004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Moodley Y, Westbury MV, Russo I-RM, Gopalakrishnan S, Rakotoarivelo A, Olsen R-A, et al. Interspecific gene flow and the evolution of specialisation in black and white rhinoceros. Mol Biol Evol. 2020;37: 3105–3117. [DOI] [PubMed] [Google Scholar]
  • 33.Li H, Durbin R. Inference of human population history from individual whole-genome sequences. Nature. 2011;475: 493–496. doi: 10.1038/nature10231 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Korneliussen TS, Albrechtsen A, Nielsen R. ANGSD: Analysis of Next Generation Sequencing Data. BMC Bioinformatics. 2014;15: 356. doi: 10.1186/s12859-014-0356-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Hudson RR. Generating samples under a Wright–Fisher neutral model of genetic variation. Bioinformatics. 2002;18: 337–338. doi: 10.1093/bioinformatics/18.2.337 [DOI] [PubMed] [Google Scholar]
  • 36.Renaud G, Hanghøj K, Korneliussen TS, Willerslev E, Orlando L. Joint Estimates of Heterozygosity and Runs of Homozygosity for Modern and Ancient Samples. Genetics. 2019;212: 587–614. doi: 10.1534/genetics.119.302057 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Kardos M, Åkesson M, Fountain T, Flagstad Ø, Liberg O, Olason P, et al. Genomic consequences of intensive inbreeding in an isolated wolf population. Nat Ecol Evol. 2018;2: 124–131. doi: 10.1038/s41559-017-0375-4 [DOI] [PubMed] [Google Scholar]
  • 38.Elferink MG, van As P, Veenendaal T, Crooijmans RPMA, Groenen MAM. Regional differences in recombination hotspots between two chicken populations. BMC Genet. 2010;11: 11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Del Priore L, Pigozzi MI. Broad-scale recombination pattern in the primitive bird Rhea americana (Ratites, Palaeognathae). PLoS One. 2017;12: e0187549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Prasad A, Lorenzen ED, Westbury MV. Evaluating the role of reference-genome phylogenetic distance on evolutionary inference. Mol Ecol Resour. 2021. doi: 10.1111/1755-0998.13457 [DOI] [PubMed] [Google Scholar]
  • 41.Liu S, Westbury MV, Dussex N, Mitchell KJ, Sinding M-HS, Heintzman PD, et al. Ancient and modern genomes unravel the evolutionary history of the rhinoceros family. Cell. 2021;184: 4874–4885.e16. doi: 10.1016/j.cell.2021.07.032 [DOI] [PubMed] [Google Scholar]
  • 42.Chikhi L, Rodríguez W, Grusea S, Santos P, Boitard S, Mazet O. The IICR (inverse instantaneous coalescence rate) as a summary of genomic diversity: insights into demographic inference and model choice. Heredity. 2018;120: 13–24. doi: 10.1038/s41437-017-0005-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Johri P, Riall K, Becher H, Excoffier L, Charlesworth B, Jensen JD. The Impact of Purifying and Background Selection on the Inference of Population History: Problems and Prospects. Mol Biol Evol. 2021;38: 2986–3003. doi: 10.1093/molbev/msab050 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.De Cahsan B, Westbury MV. Complete mitochondrial genomes offer insights into the evolutionary relationships and comparative genetic diversity of New Zealand’s iconic kiwi (Apteryx spp.). N Z J Zool. 2020; 1–9. [Google Scholar]
  • 45.Grealy A, Phillips M, Miller G, Gilbert MTP, Rouillard J-M, Lambert D, et al. Eggshell palaeogenomics: Palaeognath evolutionary history revealed through ancient nuclear and mitochondrial DNA from Madagascan elephant bird (Aepyornis sp.) eggshell. Mol Phylogenet Evol. 2017;109: 151–163. [DOI] [PubMed] [Google Scholar]
  • 46.Jetz W, Thomas GH, Joy JB, Hartmann K, Mooers AO. The global diversity of birds in space and time. Nature. 2012;491: 444–448. doi: 10.1038/nature11631 [DOI] [PubMed] [Google Scholar]
  • 47.Ottenburghs J, Ydenberg RC, Van Hooft P. The Avian Hybrids Project: gathering the scientific literature on avian hybridization. Ibis. 2015;157: 892–894. [Google Scholar]
  • 48.Price TD, Bouvier MM. The evolution of F1 postzygotic incompatibilities in birds. Evolution. 2002;56: 2083–2089. [PubMed] [Google Scholar]
  • 49.Tennyson AJD, Tomotani BM. A new fossil species of kiwi (Aves: Apterygidae) from the mid-Pleistocene of New Zealand. Hist Biol. 2021; 1–9. [Google Scholar]
  • 50.Ramstad KM, Pfunder M, Robertson HA, Colbourne RM, Allendorf FW, Daugherty CH. Fourteen microsatellite loci cross-amplify in all five kiwi species (Apteryx spp.) and reveal extremely low genetic variation in little spotted kiwi (A. owenii). Conserv Genet Resour. 2010;2: 333–336. [Google Scholar]
  • 51.Germano J, Barlow S, Castro I, Colbourne R, Cox M, Gillies C, et al. Kiwi Recovery Plan 2018–2028 Mahere Whakaora Kiwi 2018–2028. Wellington: New Zealand Department of Conservation. 2018. https://www.doc.govt.nz/globalassets/documents/science-and-technical/tsrp64entire.pdf [Google Scholar]
  • 52.Holzapfel S, Robertson HA, McLennan JA, Sporle W, Hackwell K, Impey M. Kiwi (Apteryx spp.) recovery plan. Threatened Species Recovery Plan Ser. 2008;60. http://www.doc.govt.nz/documents/science-and-technical/tsrp60.pdf [Google Scholar]
  • 53.Taylor HR, Robertson H, Carter AL, Ramstad KM. The conservation management implications of isolation by distance and high genetic diversity in Great Spotted Kiwi (Apteryx haastii). Emu—Austral Ornithology. 2021; 1–13. [Google Scholar]
  • 54.Ninkovich D. Pleistocene volcanic eruptions in New Zealand recorded in deep-sea sediments. Earth Planet Sci Lett. 1968;4: 89–102. [Google Scholar]
  • 55.Pelejero C, Calvo E, Logan GA, De Deckker P. Marine isotopic stage 5e in the southwest pacific: Similarities with Antarctica and ENSO inferences. Geophys Res Lett. 2003;30: 2185. [Google Scholar]
  • 56.Matthews NE, Smith VC, Costa A, Durant AJ, Pyle DM, Pearce NJG. Ultra-distal tephra deposits from super-eruptions: Examples from Toba, Indonesia and Taupo Volcanic Zone, New Zealand. Quat Int. 2012;258: 54–79. [Google Scholar]
  • 57.Vandergoes MJ, Hogg AG, Lowe DJ, Newnham RM, Denton GH, Southon J, et al. A revised age for the Kawakawa/Oruanui tephra, a key marker for the Last Glacial Maximum in New Zealand. Quat Sci Rev. 2013;74: 195–201. [Google Scholar]
  • 58.Hogg A, Lowe DJ, Palmer J, Boswijk G, Ramsey CB. Revised calendar date for the Taupo eruption derived by 14C wiggle-matching using a New Zealand kauri 14C calibration data set. Holocene. 2012;22: 439–449. [Google Scholar]

Decision Letter 0

Susanne P Pfeifer

29 Apr 2022

PONE-D-22-08093Genomic insights into the evolutionary relationships and demographic history of kiwiPLOS ONE

Dear Dr. Westbury,

Thank you for submitting your manuscript to PLOS ONE. After careful consideration, we feel that it has merit but does not fully meet PLOS ONE’s publication criteria as it currently stands. Therefore, we invite you to submit a revised version of the manuscript that addresses the points raised during the review process (detailed below).

Please submit your revised manuscript by Jun 13 2022 11:59PM. If you will need more time than this to complete your revisions, please reply to this message or contact the journal office at plosone@plos.org. When you're ready to submit your revision, log on to https://www.editorialmanager.com/pone/ and select the 'Submissions Needing Revision' folder to locate your manuscript file.

Please include the following items when submitting your revised manuscript:

  • A rebuttal letter that responds to each point raised by the academic editor and reviewer(s). You should upload this letter as a separate file labeled 'Response to Reviewers'.

  • A marked-up copy of your manuscript that highlights changes made to the original version. You should upload this as a separate file labeled 'Revised Manuscript with Track Changes'.

  • An unmarked version of your revised paper without tracked changes. You should upload this as a separate file labeled 'Manuscript'.

If you would like to make changes to your financial disclosure, please include your updated statement in your cover letter. Guidelines for resubmitting your figure files are available below the reviewer comments at the end of this letter.

If applicable, we recommend that you deposit your laboratory protocols in protocols.io to enhance the reproducibility of your results. Protocols.io assigns your protocol its own identifier (DOI) so that it can be cited independently in the future. For instructions see: https://journals.plos.org/plosone/s/submission-guidelines#loc-laboratory-protocols. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols.

We look forward to receiving your revised manuscript.

Kind regards,

Susanne P. Pfeifer

Academic Editor

PLOS ONE

Journal Requirements:

When submitting your revision, we need you to address these additional requirements.

1. Please ensure that your manuscript meets PLOS ONE's style requirements, including those for file naming. The PLOS ONE style templates can be found at 

https://journals.plos.org/plosone/s/file?id=wjVg/PLOSOne_formatting_sample_main_body.pdf and 

https://journals.plos.org/plosone/s/file?id=ba62/PLOSOne_formatting_sample_title_authors_affiliations.pdf

2. Thank you for stating the following in the Acknowledgments Section of your manuscript: 

"This work was supported by the Independent Research Fund Denmark | Natural Sciences, Forskningsprojekt 1, grant no. 8021-00218B, and the Villum Fonden Young Investigator Programme, grant no. 13151, to EDL. We would like to acknowledge the Te Parawhau Trust and Waikaremoana iwi, Te Rūnanga o Ngāi Tahu, Te Ātiawa Manawhenua Ki Te Tau Ihu Trust, who provided guidance to the authors who generated the kiwi genomic data, on which our study is based. The authors declare no conflicts of interest."

We note that you have provided funding information that is not currently declared in your Funding Statement. However, funding information should not appear in the Acknowledgments section or other areas of your manuscript. We will only publish funding information present in the Funding Statement section of the online submission form. 

Please remove any funding-related text from the manuscript and let us know how you would like to update your Funding Statement. Currently, your Funding Statement reads as follows: 

"This work was supported by the Independent Research Fund Denmark | Natural Sciences, Forskningsprojekt 1, grant no. 8021-00218B, and the Villum Fonden Young Investigator Programme, grant no. 13151, to EDL. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript."

Please include your amended statements within your cover letter; we will change the online submission form on your behalf.

Additional Editor Comments:

I have now received reviews from two experts in the field and their reading matches my own. Namely, given that this manuscript is based on publicly available genomic data, the authors will need to put their work in the appropriate context to demonstrate their novel contributions beyond the previously published work. Each reviewer makes additional points of importance with regards to the methodology as well as the interpretation of the results that would also need to be fully addressed. Of particular importance, the authors will need to provide evidence that the inferred model actually matches their data (and that PSMC is able to recapitulate the observations made from the data simulated under the inferred model). In this regard, the authors need to be mindful about the fact that their dataset contains regions affected by direct and linked selection, which can result in serious mis-inference of the population's demographic history (see Johri et al. MBE 2021). To allow the authors to address these comments / concerns, I recommend a major revision.

Reviewers' comments:

Reviewer's Responses to Questions

Comments to the Author

1. Is the manuscript technically sound, and do the data support the conclusions?

The manuscript must describe a technically sound piece of scientific research with data that supports the conclusions. Experiments must have been conducted rigorously, with appropriate controls, replication, and sample sizes. The conclusions must be drawn appropriately based on the data presented.

Reviewer #1: Partly

Reviewer #2: Partly

**********

2. Has the statistical analysis been performed appropriately and rigorously?

Reviewer #1: No

Reviewer #2: Yes

**********

3. Have the authors made all data underlying the findings in their manuscript fully available?

The PLOS Data policy requires authors to make all data underlying the findings described in their manuscript fully available without restriction, with rare exception (please refer to the Data Availability Statement in the manuscript PDF file). The data should be provided as part of the manuscript or its supporting information, or deposited to a public repository. For example, in addition to summary statistics, the data points behind means, medians and variance measures should be available. If there are restrictions on publicly sharing data—e.g. participant privacy or use of data from a third party—those must be specified.

Reviewer #1: Yes

Reviewer #2: Yes

**********

4. Is the manuscript presented in an intelligible fashion and written in standard English?

PLOS ONE does not copyedit accepted manuscripts, so the language in submitted articles must be clear, correct, and unambiguous. Any typographical or grammatical errors should be corrected at revision, so please note any specific errors here.

Reviewer #1: Yes

Reviewer #2: Yes

**********

5. Review Comments to the Author

Please use the space provided to explain your answers to the questions above. You may also include additional comments for the author, including concerns about dual publication, research ethics, or publication ethics. (Please upload your review as an attachment if it exceeds 20,000 characters)

Reviewer #1: This project uses previously published genomic data from five species of kiwi birds (and their close relatives, as outgroups) to understand the evolutionary history of this species. The project uses a set of population genomic methods to infer divergence times, introgression history, and inbreeding.

This manuscript has many strengths.

1. I like that the manuscript has a tight focus. While I always enjoy reading a paper that has broad implications, it is also always really nice to read a paper that focuses on the system and the data. It is nicely concise.

2. The study very carefully handles reference bias, which few comparative genomic projects even consider.

3. The figures are generally clear and easy to follow.

I have a few major suggestions for improvement.

1. There have been two really comprehensive papers published on kiwi evolutionary dynamics already, both out of Jason Weir's research group. This study cites both papers and references them a few times, but I found more should be done here to (1) explain in the introduction how this study can add to these existing studies and (2) compare results found across studies in the Discussion. Especially because the discussion is so focused on the kiwi clade itself, it should more strongly integrate what we already knew about kiwi diversification.

2. The study uses hPSMC (as I understand it) to infer gene flow between the species. I have never used hPSMC, but I have used PSMC, so I was curious about this application. I read the manual, and I cannot find any evidence that hPSMC allows inference of gene flow. I would recommend the authors either (1) clarify how they modified hPSMC to allow inference of gene flow or (2) clarify the language so that it no longer seems like hPSMC was used for gene flow inference. The discrepancies between divergence times inferred from a phylogeny could be due to ILS, post-divergence gene flow and / or methodological uncertainty. Given the QuIBL results, ILS and / or methodological uncertainty seems most likely.

3. The study often refers to the importance of ensuring that the reference genome used doesn't introduce bias. Given that this is a recurrent theme in the manuscript, I would suggest discussing this aspect of the study in the introduction and / or discussion to summarize what is known about this issue and why it matters.

MINOR COMMENTS

- L19 - 20: I might be over-interpreting this sentence, but fossils are unlikely to help infer evolutionary relationships -- just evolutionary timings. (See again L61 - 63.)

- L58 - 60: "In contrast" -- what are the authors contrasting here?

- L82: the Bemmels et al 2021 study also showed this, using whole genome data.

- L101 - 104: Recommend listing species names here again.

- L135: Consider reporting the total number of gene trees inferred.

- L154 - 156: What is the justification for these BIC thresholds?

- L159 - 169: Consider expanding the discussion of the hPSMC approach here. Most readers will be familiar with the PSMC approach, but might not fully appreciate the difference between PSMC and hPSMC.

- L171: What is inferred here is technically the substitution rate, no?

- L199: For this analysis, which variant set was used? The one inferred using ANGSD (L130)?

- L236: Recommend striking "earliest to diverge" here; see this great blog post here for why: http://for-the-love-of-trees.blogspot.com/2016/09/the-ancestors-are-not-among-us.html

- L237 - 238: 11% and 26% actually not that high! I would recommend reviewing some other phylogenomic discordance papers.

- L241: To my interpretation, the QuIBL results very strongly indicated this is likely just ILS.

- L243 - 253: I find this analysis a bit confusing. The study is comparing divergence times inferred from the phylogeny (which, it would be good to remind the readers here of how these were inferred and how deep they were) to timings inferred from hPSMC? We would expect these two estimates to be very different because they are measuring very different aspects of divergence history. This paragraph makes it clear that this discrepancy could be because of ILS or post-divergence gene flow. Yet, the first sentence of this paragraph seems to suggest this analysis was done to tell us about post-divergence gene flow, but this approach isn't the best approach to study post-divergence gene flow. (Also, another possibility for why these estimates are off: the difficulties in time calibration given uncertainties in substitution rate, fossil dates, and generation time. See also L318 - 320 for the same.) The study appears to be more conclusive that it is post-divergence gene flow in L431, which confuses me.

- L273 - 278: I would consider striking data on the other palaeognath species as they aren't really part of the paper otherwise.

- L304 - 306: Can the authors clarify how the hPSMC results show that there was continual gene flow? The high levels of discordance at the site level (Table S6) are in-line with what has been seen in other phylogenomic studies and could simply be the result of ILS.

- L322 - L323: High dispersal in birds is one hypothesis for why there is more hybridization, but another hypothesis is that birds seem to evolve postzygotic isolation slowly (Price an Bouvier 2002)

- L327 - L329: One aspect that confuses me here - these species seem to have abutting / overlapping ranges in the present day. Also current ranges do not necessarily reflect historical ranges. So, it is quite possible these species were co-occurring, which would have allowed hybridization pretty easily. And, even with no flight, these birds still disperse? Or is dispersal essentially non-existent?

- L361 - 3: Here is an example where the Bemmels paper should definitely be cited and used as a comparison point, as they conducted a more formal analysis of the same question.

- L422 - 428: Here is another place where the previous demographic work on kiwi should be cited.

- Figure 1: What is the citation for the historical ranges for these species?

Reviewer #2: In this study, Westbury et al. analyze representative whole genomes from each of five kiwi species to investigate the evolutionary history of this group. They primarily assessed phylogenetic relationships and looked for evidence of phylogenetic discordance due to gene flow and/or ILS, and they inferred historical effective population sizes with PSMC. The authors recover a phylogeny consistent with prior studies, and find some evidence of gene flow and/or ILS, although these latter results are not well resolved. Further, the authors find overall levels of genome-wide diversity that appear consistent with the effective population size histories (Ne of ~25k or less over the past million years), and some ROH, potentially indicative of inbreeding, in the little spotted kiwi.

The study exclusively uses data generated from other studies, which also evaluated the evolution and origins of kiwi. However, these prior studies are not sufficiently acknowledged in this manuscript. The data are not tied to their original sources (Sackton et al. 2019, Weir et al. 2016, Bemmels et al. 2021, more?), and the major findings and unresolved issues from the prior studies are not provided. Further, some important details from the methods are missing, and some of the results appear not to support one another (i.e. the ILS and gene flow analyses). Lastly, some of the conclusions based on the Ne trajectories relating to environmental shifts are unsupported. Given the work that has been done before, I am not sure what the contribution of this study is. In terms of its scope and aims, it is not well differentiated from the previous studies, which are probably more suited to investigating the evolutionary history of this system because they included more data. Overall, the manuscript was relatively clear and the analyses are mostly appropriate, but there are a number of crucial areas that require improvement. What follows is a detailed breakdown of my concerns.

89: The summary of prior genomic studies of kiwi in this introduction is insufficient. As written, the introduction fails to acknowledge that there are a few genomic studies of kiwi already in existence, which also tackle the issues of species origin, population structure, diversity, phylogenetic discordance, etc. with far more samples (e.g. Sackton et al. 2019, Weir et al. 2016, Bemmels et al. 2021). The major findings of those studies and the remaining gaps in knowledge should be presented in the introduction. How the present study addresses those gaps should be stated. How the results from this study agree with, add to, or refute prior studies must also be presented in the Discussion.

101: Citations are needed for all sequence data and reference genomes used in this study. This information should be added to the methods here and also included in Table S1. Also, in Table S1 the accession codes for the little spotted kiwi are duplicated - typo?

115: What is the depth of sequencing coverage? This information is essential for assessing the robustness of the methods and results.

132: How many sequences were used in the phylogenetic inference? Report sample sizes for all analyses (i.e. for the ILS/gene flow tests as well).

135: For clarity, specify that the "gene trees" are actually from randomly sampled sequences across the genome, not genes.

154: Explain the rationale of this BIC cutoff. Is it based on the original QuIBL paper, or used here only?

169, 224: Provide complete details of the PSMC pipeline run parameters.

172: Given the apparent signal of some ILS and gene flow, is this a robust way to estimate mutation rate? Presumably these factors could cause the mutation rate to be under-estimated. Provide a rationale or acknowledgement, and, if possible, compare with other mutation rate estimates from kiwi or similar species.

204: Are the kiwi genomes highly contiguous? How does contiguity of the assemblies potentially impact ROH identification? Add more information to the text or supplement about the contiguity of the genome assemblies.

209: Clarify if the values in Fig. 2B were obtained from ROHan with ROH length set to 1 Mb for all species.

220: What is meant by "this species"? This part is unclear.

225: Clarify that the PSMC results were checked for overfitting (did all intervals contain >=10 recombinations, as recommended by https://github.com/lh3/psmc?).

237-241: A graphical representation of the gene flow and demographic history results should be provided, if possible. It's hard to grasp these results from the text only. The authors could look to Edelman et al. 2019 for inspiration. More detailed demographic model figures are standard in the literature, and very helpful for understanding the results, but other figures to visually show the amount of gene flow/ILS could also be used.

264-271: Is the aim to look at "recent" inbreeding? How recent? Be explicit. Length of ROH corresponds with the number of generations back to the shared common ancestor. ROH <=1 Mb can originate from shared ancestors many generations ago, which may or may not be considered "recent" depending on the context, and might not even be reflective of "inbreeding" per se, but rather of limited population size. Given that the other demographic analyses deal with far more ancient history tens to hundreds of thousands of years ago, while some of the species declines occurred far more recently in the post-colonial era within the past thousand years, the communication of these results needs to be more precise.

273: It would be helpful to add the other species to this figure.

216-22 and 281-286: Why is there a "bias" for these particular genomes when aligning to a conspecific reference? Does it have to do with the same/different individual being used for the resequencing and the assembly? This is unclear. Explain.

303-308: These results are not consistent. The QuIBL analysis suggests there is little evidence of gene tree discordance overall, consistent with some ILS, but the hPSMC results suggest continuous gene flow. Later, "long-term" ILS is mentioned as a possibility (line 319), which would be surprising given the relatively limited population sizes of these species (~20k or less under the assumed mutation rate). A more coherent explanation of the results is needed in order to tie them together and support the study conclusions. I might also suggest using additional methods to address these issues, since the analyses included in this paper do not seem sufficient or convincing.

360-361, 365: Species with long generation times will not necessarily show signals of recent bottlenecks in genome-wide heterozygosity. This study does not adequately resolve whether the observed ROH are due to very recent inbreeding or more distant common ancestry. These levels of genome-wide diversity in the kiwi genomes appear largely consistent with the long-term Ne values in each species, which is not surprising. These points need to be made clear.

414-421: It is well known that PSMC provides effectively no resolution for this time period (25 kya to present). Furthermore, inspection of Fig. 3 does not appear to show a marked and consistent signal of decline among all species. This is mostly speculation and should either be rephrased to avoid misleading the reader, or removed.

435-437: The results presented in this study are insufficient to support this claim. As stated above, there is not enough power in the analyses of these single genomes to demonstrate the impact of volcanism over other forces. Furthermore, linking PSMC curves to environmental shifts is dubious, since PSMC analyses involve several over-simplifications and assumptions. For example, the results are highly dependent on the assumed mutation rate and generation time, which are often unknown (in this case, we don't know how reliable the mutation rate is).

**********

6. PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #1: No

Reviewer #2: No

While revising your submission, please upload your figure files to the Preflight Analysis and Conversion Engine (PACE) digital diagnostic tool, https://pacev2.apexcovantage.com/. PACE helps ensure that figures meet PLOS requirements. To use PACE, you must first register as a user. Registration is free. Then, login and navigate to the UPLOAD tab, where you will find detailed instructions on how to use the tool. If you encounter any issues or have any questions when using PACE, please email PLOS at figures@plos.org. Please note that Supporting Information files do not need this step.

Decision Letter 1

Susanne P Pfeifer

15 Jul 2022

PONE-D-22-08093R1Genomic insights into the evolutionary relationships and demographic history of kiwiPLOS ONE

Dear Dr. Westbury,

Thank you for submitting your manuscript to PLOS ONE. After careful consideration, we feel that it has merit but does not fully meet PLOS ONE’s publication criteria as it currently stands. Therefore, we invite you to submit a revised version of the manuscript that addresses the points raised during the review process as detailed below.

Please submit your revised manuscript by Aug 29 2022 11:59PM. If you will need more time than this to complete your revisions, please reply to this message or contact the journal office at plosone@plos.org. When you're ready to submit your revision, log on to https://www.editorialmanager.com/pone/ and select the 'Submissions Needing Revision' folder to locate your manuscript file.

Please include the following items when submitting your revised manuscript:

  • A rebuttal letter that responds to each point raised by the academic editor and reviewer(s). You should upload this letter as a separate file labeled 'Response to Reviewers'.

  • A marked-up copy of your manuscript that highlights changes made to the original version. You should upload this as a separate file labeled 'Revised Manuscript with Track Changes'.

  • An unmarked version of your revised paper without tracked changes. You should upload this as a separate file labeled 'Manuscript'.

If you would like to make changes to your financial disclosure, please include your updated statement in your cover letter. Guidelines for resubmitting your figure files are available below the reviewer comments at the end of this letter.

If applicable, we recommend that you deposit your laboratory protocols in protocols.io to enhance the reproducibility of your results. Protocols.io assigns your protocol its own identifier (DOI) so that it can be cited independently in the future. For instructions see: https://journals.plos.org/plosone/s/submission-guidelines#loc-laboratory-protocols. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols.

We look forward to receiving your revised manuscript.

Kind regards,

Susanne P. Pfeifer

Academic Editor

PLOS ONE

Additional Editor Comments:

There remains a major concern about the reliable distinction of incomplete lineage sorting from post-divergence gene flow in this study. Currently, the authors favor one process over the other, without much convincing evidence. Additional analyses will be needed to address this issue (as well as several smaller ones highlighted by the two reviewers). Both reviewers agree that the writing itself needs to be improved as well, in particular with regards to the distinction between previously published results and novel insights from this study (which was already pointed out in the first round of reviews) and the discussion (which needs to be more focused). For reproducibility, a permanent repository for any scripts used in the analyses should be included in a future revision. 

[Note: HTML markup is below. Please do not edit.]

Reviewers' comments:

Reviewer's Responses to Questions

Comments to the Author

1. If the authors have adequately addressed your comments raised in a previous round of review and you feel that this manuscript is now acceptable for publication, you may indicate that here to bypass the “Comments to the Author” section, enter your conflict of interest statement in the “Confidential to Editor” section, and submit your "Accept" recommendation.

Reviewer #2: (No Response)

Reviewer #3: (No Response)

**********

2. Is the manuscript technically sound, and do the data support the conclusions?

The manuscript must describe a technically sound piece of scientific research with data that supports the conclusions. Experiments must have been conducted rigorously, with appropriate controls, replication, and sample sizes. The conclusions must be drawn appropriately based on the data presented.

Reviewer #2: Partly

Reviewer #3: Partly

**********

3. Has the statistical analysis been performed appropriately and rigorously?

Reviewer #2: No

Reviewer #3: Yes

**********

4. Have the authors made all data underlying the findings in their manuscript fully available?

The PLOS Data policy requires authors to make all data underlying the findings described in their manuscript fully available without restriction, with rare exception (please refer to the Data Availability Statement in the manuscript PDF file). The data should be provided as part of the manuscript or its supporting information, or deposited to a public repository. For example, in addition to summary statistics, the data points behind means, medians and variance measures should be available. If there are restrictions on publicly sharing data—e.g. participant privacy or use of data from a third party—those must be specified.

Reviewer #2: Yes

Reviewer #3: No

**********

5. Is the manuscript presented in an intelligible fashion and written in standard English?

PLOS ONE does not copyedit accepted manuscripts, so the language in submitted articles must be clear, correct, and unambiguous. Any typographical or grammatical errors should be corrected at revision, so please note any specific errors here.

Reviewer #2: Yes

Reviewer #3: Yes

**********

6. Review Comments to the Author

Please use the space provided to explain your answers to the questions above. You may also include additional comments for the author, including concerns about dual publication, research ethics, or publication ethics. (Please upload your review as an attachment if it exceeds 20,000 characters)

Reviewer #2: I have read the revised manuscript from Westbury et al. and find that although it has improved, there are outstanding issues. I am mostly satisfied that the study has been differentiated from already published works. However, my biggest concern is that the conclusions about incomplete lineage sorting (ILS) versus post-divergence gene flow are not supported by the evidence. Also, a couple of minor issues I mentioned in my original review were not adequately addressed. See below for complete details. I recommend further revision before the manuscript can be accepted for publication.

ILS/gene flow analyses: The presentation of the ILS/gene flow results are slightly more clear now, and after some further thought, I am not convinced that the results favor post-divergence gene flow over ILS. First, hPSMC simply does not distinguish between these two processes, so it is not sufficient evidence either way. Second, QuIBL is supposed to specifically differentiate between ILS+gene flow and ILS alone, and the authors simply chose to disagree with the QuIBL results tending to support the ILS alone model. The argument of site discordance within congruent windows suggesting post-divergence gene flow does not make sense to me. What seems to be happening is that kiwi diverged very recently, and may really be incipient species or even subspecies that are still capable of hybridizing. Kiwi have long generation times, apparently only diverged within the last few million years, and it is well known that bird genomes evolve slowly. Under this scenario, I would expect ILS to be more of a factor than post-divergence gene flow. In my original review, I suggested performing additional analyses to better resolve this question. I reiterate that suggestion here. One possibility might be to use the ABBA/BABA test, but there are probably other options as well.

Discussion: At about four pages, the Discussion is far too long. Compare with the results, which are just over one page in length. The Discussion needs to be condensed, and the main conclusions that are supported by the analyses in the paper should be stated clearly. As is, the Discussion is meandering, not very well organized, and hard to digest.

115-123: Briefly mention the mean or range of sequencing depths.

165: I don't advocate for using the term "locus trees" instead of "gene trees". Gene tree is the terminology used in the literature. It would be better to simply state, once, that the "gene trees" are actually just from windows, not genes.

Table S1: The previously published data (reads, assemblies) should have citations for sources. A supplemental table being too "busy" isn't really justification for failing to give proper attribution and accurately document where the data came from.

Reviewer #3: In this manuscript, the authors re-examine the evolutionary relationships between kiwi species and look for introgression between these species. I like the introduction as well as the figures and see value in publishing confirmatory results besides clear values of nucleotide diversity (for instance in Figure 2). However, as most of their results are confirmatory, which results are confirmatory and what has been gained from performing this study should be made explicitly clear. Here are my more specific comments:

1. Please provide a link (at Github or Dryad etc) to the scripts used to perform the analyses in this study.

2. In the abstract, the authors claim that they uncover similarities and differences in the demographic histories of the kiwi species studied here. However, as Weir et al (2016) already uncovers similarities in the demographic histories using PSMC, the authors should mention that they confirm previous observations of past population size changes. If the authors disagree with this, then they should mention in the abstract what they uncovered that is different.

3. Overall, I really like Figure 2. Could you perhaps also provide the synonymous site heterozygosity or heterozygosity calculated only using 4-fold degenerate sites as well? I suggest that as it may be helpful later for other studies.

4. Lines 347-360: We now know from multiple studies that unaccounted for factors such as the presence of purifying selection across the genome (Johri et al 2021) and hidden substructure in populations (Chikhi et al 2018) can result in biased inference of population history when using MSMC/ PSMC -like approaches. Although there has been no thorough investigation of how hPSMC would be biased by the same factors, it most likely would. Thus, the authors should discuss how this could affect their results.

The same reasons might be responsible for observing similar temporal patterns of population decline across all five species, as mentioned in lines 491-493.

5. This is simply a suggestion, and the authors can decide whether they like it or not. You could place the section about mapping to different reference genomes and obtaining the same species tree in the Results section (instead of the Methods section), if you like, as I believe that it is an interesting result. However, this is entirely up to the authors.

6. The discussion section currently meanders a bit. Subsections with headings will really help the reader get the main points from the discussion.

Minor Comments:

1. Lines 368-370: could you please refer to the respective figures from your manuscript.

References:

Chikhi L et al. 2018. The IICR (inverse instantaneous coalescence rate) as a summary of genomic diversity: insights into demographic inference and model choice. Heredity. 120:13–24.

Johri P et al. 2021. The impact of purifying and background selection on the inference of population history: problems and prospects. Mol. Biol. Evol. 38:2986–3003.

**********

7. PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #2: No

Reviewer #3: No

**********

[NOTE: If reviewer comments were submitted as an attachment file, they will be attached to this email and accessible via the submission site. Please log into your account, locate the manuscript record, and check for the action link "View Attachments". If this link does not appear, there are no attachment files.]

While revising your submission, please upload your figure files to the Preflight Analysis and Conversion Engine (PACE) digital diagnostic tool, https://pacev2.apexcovantage.com/. PACE helps ensure that figures meet PLOS requirements. To use PACE, you must first register as a user. Registration is free. Then, login and navigate to the UPLOAD tab, where you will find detailed instructions on how to use the tool. If you encounter any issues or have any questions when using PACE, please email PLOS at figures@plos.org. Please note that Supporting Information files do not need this step.

Decision Letter 2

Susanne P Pfeifer

20 Sep 2022

PONE-D-22-08093R2Genomic insights into the evolutionary relationships and demographic history of kiwiPLOS ONE

Dear Dr. Westbury,

Thank you for submitting a revised version of your manuscript to PLOS ONE. Most of the reviewers' comments were addressed in this revision, however there remain a few outstanding points that should be addressed: 1) minor modifications in the text would help to improve clarity and 2) p-values should be included in Table S9 to strengthen the analysis.

Please submit your revised manuscript by Nov 04 2022 11:59PM. If you will need more time than this to complete your revisions, please reply to this message or contact the journal office at plosone@plos.org. When you're ready to submit your revision, log on to https://www.editorialmanager.com/pone/ and select the 'Submissions Needing Revision' folder to locate your manuscript file.

Please include the following items when submitting your revised manuscript:

  • A rebuttal letter that responds to each point raised by the academic editor and reviewer(s). You should upload this letter as a separate file labeled 'Response to Reviewers'.

  • A marked-up copy of your manuscript that highlights changes made to the original version. You should upload this as a separate file labeled 'Revised Manuscript with Track Changes'.

  • An unmarked version of your revised paper without tracked changes. You should upload this as a separate file labeled 'Manuscript'.

If you would like to make changes to your financial disclosure, please include your updated statement in your cover letter. Guidelines for resubmitting your figure files are available below the reviewer comments at the end of this letter.

If applicable, we recommend that you deposit your laboratory protocols in protocols.io to enhance the reproducibility of your results. Protocols.io assigns your protocol its own identifier (DOI) so that it can be cited independently in the future. For instructions see: https://journals.plos.org/plosone/s/submission-guidelines#loc-laboratory-protocols. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols.

We look forward to receiving your revised manuscript.

Kind regards,

Susanne P. Pfeifer

Academic Editor

PLOS ONE

Journal Requirements:

Please review your reference list to ensure that it is complete and correct. If you have cited papers that have been retracted, please include the rationale for doing so in the manuscript text, or remove these references and replace them with relevant current references. Any changes to the reference list should be mentioned in the rebuttal letter that accompanies your revised manuscript. If you need to cite a retracted article, indicate the article’s retracted status in the References list and also include a citation and full reference for the retraction notice.

Additional Editor Comments (if provided):

[Note: HTML markup is below. Please do not edit.]

Reviewers' comments:

Reviewer's Responses to Questions

Comments to the Author

1. If the authors have adequately addressed your comments raised in a previous round of review and you feel that this manuscript is now acceptable for publication, you may indicate that here to bypass the “Comments to the Author” section, enter your conflict of interest statement in the “Confidential to Editor” section, and submit your "Accept" recommendation.

Reviewer #2: (No Response)

Reviewer #3: (No Response)

********** 

2. Is the manuscript technically sound, and do the data support the conclusions?

The manuscript must describe a technically sound piece of scientific research with data that supports the conclusions. Experiments must have been conducted rigorously, with appropriate controls, replication, and sample sizes. The conclusions must be drawn appropriately based on the data presented.

Reviewer #2: Yes

Reviewer #3: (No Response)

********** 

3. Has the statistical analysis been performed appropriately and rigorously?

Reviewer #2: Yes

Reviewer #3: (No Response)

********** 

4. Have the authors made all data underlying the findings in their manuscript fully available?

The PLOS Data policy requires authors to make all data underlying the findings described in their manuscript fully available without restriction, with rare exception (please refer to the Data Availability Statement in the manuscript PDF file). The data should be provided as part of the manuscript or its supporting information, or deposited to a public repository. For example, in addition to summary statistics, the data points behind means, medians and variance measures should be available. If there are restrictions on publicly sharing data—e.g. participant privacy or use of data from a third party—those must be specified.

Reviewer #2: Yes

Reviewer #3: No

********** 

5. Is the manuscript presented in an intelligible fashion and written in standard English?

PLOS ONE does not copyedit accepted manuscripts, so the language in submitted articles must be clear, correct, and unambiguous. Any typographical or grammatical errors should be corrected at revision, so please note any specific errors here.

Reviewer #2: Yes

Reviewer #3: (No Response)

********** 

6. Review Comments to the Author

Please use the space provided to explain your answers to the questions above. You may also include additional comments for the author, including concerns about dual publication, research ethics, or publication ethics. (Please upload your review as an attachment if it exceeds 20,000 characters)

Reviewer #2: I thank the authors for addressing my previous comments. I have a few more relatively minor concerns with regard to some of the changes in the revised manuscript, as detailed below.

26-27: Would help to state which species this is referring to, or what the exceptions are.

188: Define fb

189-190: Blocks of what size were used?

307-311: These sentences are not sufficiently clear: 307: Similarities between what? the ingroup and outgroup? Is the bias in the emu's diversity, or the kiwi's? 310: Why is the bias caused by using a conspecific reference as opposed to a heterospecific reference? You don't actually know which produces the more accurate result.

310: Fewer not less

412-423: I suggest removing this part of the Discussion. It's very hard to comprehend and is better understood by visual by examination of the figures. Obviously the specifics are complicated and there are signals of gene flow/ILS throughout the phylogeny, so I recommend just summarizing with the big picture: there are signals of gene flow and/or ILS throughout the tree, which is not too surprising for a very young radiation like this. I still think the analyses of gene flow v. ILS are a little shaky, and deeper investigations of particular species histories should be left to future studies with larger datasets (ie, more individuals).

Fig 2: All cells of the matrix are a uniform shade of grey, so the results cannot be seen. Also, what is the basis for the gene flow events "hypothesized" by the authors? Explain this, or, preferably, eliminate and just show the actual results.

Table S3: Define N50, L75

Table S8: "production" should be "proportion" ?

Table S9: I have not used this method of calculating f-branch statistics, but the fact that no p-values could be computed for about half the table is very disconcerting. What is the explanation? Have the authors made an effort to rectify this? I consider this a serious flaw with this analysis.

Reviewer #3: The authors have addressed most of the comments.

The github link is not really working and so the data availability could not be verified.

********** 

7. PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #2: No

Reviewer #3: No

**********

[NOTE: If reviewer comments were submitted as an attachment file, they will be attached to this email and accessible via the submission site. Please log into your account, locate the manuscript record, and check for the action link "View Attachments". If this link does not appear, there are no attachment files.]

While revising your submission, please upload your figure files to the Preflight Analysis and Conversion Engine (PACE) digital diagnostic tool, https://pacev2.apexcovantage.com/. PACE helps ensure that figures meet PLOS requirements. To use PACE, you must first register as a user. Registration is free. Then, login and navigate to the UPLOAD tab, where you will find detailed instructions on how to use the tool. If you encounter any issues or have any questions when using PACE, please email PLOS at figures@plos.org. Please note that Supporting Information files do not need this step.

Decision Letter 3

Susanne P Pfeifer

28 Sep 2022

Genomic insights into the evolutionary relationships and demographic history of kiwi

PONE-D-22-08093R3

Dear Dr. Westbury,

We’re pleased to inform you that your manuscript has been judged scientifically suitable for publication and will be formally accepted for publication once it meets all outstanding technical requirements.

Within one week, you’ll receive an e-mail detailing the required amendments. When these have been addressed, you’ll receive a formal acceptance letter and your manuscript will be scheduled for publication.

An invoice for payment will follow shortly after the formal acceptance. To ensure an efficient process, please log into Editorial Manager at http://www.editorialmanager.com/pone/, click the 'Update My Information' link at the top of the page, and double check that your user information is up-to-date. If you have any billing related questions, please contact our Author Billing department directly at authorbilling@plos.org.

If your institution or institutions have a press office, please notify them about your upcoming paper to help maximize its impact. If they’ll be preparing press materials, please inform our press team as soon as possible -- no later than 48 hours after receiving the formal acceptance. Your manuscript will remain under strict press embargo until 2 pm Eastern Time on the date of publication. For more information, please contact onepress@plos.org.

Kind regards,

Susanne P. Pfeifer

Academic Editor

PLOS ONE

Acceptance letter

Susanne P Pfeifer

30 Sep 2022

PONE-D-22-08093R3

Genomic insights into the evolutionary relationships and demographic history of kiwi

Dear Dr. Westbury:

I'm pleased to inform you that your manuscript has been deemed suitable for publication in PLOS ONE. Congratulations! Your manuscript is now with our production department.

If your institution or institutions have a press office, please let them know about your upcoming paper now to help maximize its impact. If they'll be preparing press materials, please inform our press team within the next 48 hours. Your manuscript will remain under strict press embargo until 2 pm Eastern Time on the date of publication. For more information please contact onepress@plos.org.

If we can help with anything else, please email us at plosone@plos.org.

Thank you for submitting your work to PLOS ONE and supporting open access.

Kind regards,

PLOS ONE Editorial Office Staff

on behalf of

Dr. Susanne P. Pfeifer

Academic Editor

PLOS ONE

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Supplementary Materials

    S1 File. File containing all supplementary information: S1-S9 Tables and S1-S5 Figs.

    (DOCX)

    Attachment

    Submitted filename: Response to reviewers.docx

    Attachment

    Submitted filename: Reviewer responses.docx

    Attachment

    Submitted filename: Reviewer responses.docx

    Data Availability Statement

    All raw data used in this manuscript were previously published and the accession codes are listed in supplementary table S2. Example commands and the scripts necessary to perform all analyses in this manuscript are available at github.com/Mvwestbury/Kiwi-genomes.


    Articles from PLoS ONE are provided here courtesy of PLOS

    RESOURCES