Background

Epigenetics plays a critical role in regulating the expression of genes during embryonic development and tissue differentiation in living organisms. The epigenome is defined as the structural adaptation of chromosomal regions [1], which maintains stable cellular processes in living individuals unless disrupted by external factors. The epigenetic modifications involve complex mechanisms governed by many structural and molecular regulators. The common mechanisms include (a) DNA methylation, (b) histone protein modification, and (c) non-coding RNAs [2]. Among all epigenetic marks, DNA methylation has been the most extensively studied [3]. The most common structural modification is the addition of a methyl group to the 5th carbon atom of a cytosine base, forming 5-methylcytosine (5-mC). The 5-mC modification in a DNA segment changes gene activity without changing the gene sequence [2, 4]. Hypermethylation in regulatory regions, such as promoters in the genome, is known to suppress transcription [5, 6]. In contrast, the loss of methylation marks, leading to hypomethylation, stimulates the expression of certain genes [7]. A methyl group-bound cytosine base is typically followed by guanine base in the sequence; this short segment is known as a CpG dinucleotide. In vertebrates, methylation most commonly occurs in the CpG context [8]. The percentage of methylated CpG sites varies by cell type [9] and individual organism. Any minor change to the epigenome can exert an impact on cellular function [10].

Surrogate production involves advanced reproductive biology-based biotechnology, which enables the production of donor-derived gametes via germline chimeras. Isolated germ stem cells (GSCs) from a donor species are implanted into the suitable sterile recipient (endogenous germ-cell-free individuals) to produce donor-derived gametes [11]. Surrogate production technologies have offered the possibility of preserving the germplasm of large-bodied commercially valuable species in smaller species with a shorter sexual maturation cycle [12, 13]. This technique provides a unique opportunity to cryopreserve maternal genetic information in fish, as neither fish eggs nor embryos can be cryopreserved. Therefore, there are efforts to develop techniques to conserve endangered species such as sturgeon [14,15,16]. Despite the recent advancements in this field, the practical application of this technology in aquaculture production is still questionable. Our understanding is limited to the morphological characteristics of the gametes and progeny produced via surrogacy. There are several questions pertaining to the reliability of this approach must be answered to maximize the possibility of applying surrogate production technologies in aquaculture. Epigenetic change is a major question associated with germ cell transplantation (GCT) that has never been reported. This mechanism is sensitive to internal and external environmental factors and can induce heritable phenotypical alterations [17]. Our knowledge of how and to what degree GCT can modify the epigenome (DNA methylation) of the donor germline is still scarce. Elucidating the epigenetic perturbations and their functional consequences will provide us insight into the reliability of the surrogate production approach.

Several factors may have a significant impact on the donor’s methylome during GCT. Donor GSC is subjected to systematic procedures and is exposed to a variety of chemical compounds during surrogate production. The complex gonadal tissue is then reduced to single cells and transplanted into recipients. Cell proliferation and differentiation of donor GSCs occur in the recipient’s gonad after transplantation, supported by the gonadal somatic cells of the recipient. Very few of the hundreds of transplanted cells survive to be incorporated into the recipient gonad. In a previous study, Saito et al. [18] reported the size difference of the larvae produced by surrogate in zebrafish. Franek et al. [13] observed different morphological characteristics in male and female donor-derived gametes. Therefore, we hypothesize that the methylome of donor GSCs may undergo alterations in the recipient gonad and may thus lose or gain some crucial methylation marks. Considering these reports, it is now crucial to think beyond the morphological characteristics of the gametes produced by the surrogate parent. The number of methylation studies on donor-derived gametes produced via GCT is considerably lower than any other epigenetic study. A study of spermatogonia transplantation in mice has been reported, and it described the methylation of donor-derived spermatozoa and their offspring, but the results were limited to only a few imprinting genes [19]. In vitro culturing and transplantation of foetal germ cells have been conducted in mice, and the derived offspring were reported to exhibit growth abnormalities due to differential methylation [20], whereas fresh testicular tissue grafting into the recipients led to no epigenetic abnormalities in the germ cells or the offspring [21]. However, no information on methylation alterations in donor-derived gametes in fish has been reported.

It was previously reported that in zebrafish (Danio rerio), DNA methylation pattern was preserved in the germline [22], and the paternal methylome was inherited by the embryos [23, 24], which is why it is crucial to decode the methylome of sperm produced by the surrogate. It has been shown that after endogenous germ cell depletion, all sterile zebrafish transdifferentiated into males [25, 26] and there was no female germline chimera after transplantation. Therefore, we used donor-derived sperm samples and progeny for this study. Sterilization to deplete the endogenous germ cells of the host species is a crucial step in surrogate production to prevent the production of recipient-derived gametes. Sterilization of a host can be achieved by chromosomal manipulation [27], gene knockdown, gene knockout [28], UV exposure [29], and chemical treatment [30]. In this study, we used vas::EGFP transgenic zebrafish as donors and two types of hosts: (1) dead end (dnd) gene knockdown (dnd gene is exclusively expressed in primordial germ cells (PGCs) and plays a crucial role in PGC migration to the gonadal ridge during embryonic development, dnd-morpholino (dnd-MO) prevents PGC migration by blocking the translation of dead end protein [31]; hence the recipient becomes sterile with no endogenous germ cells) and (2) zebrafish and pearl danio (Danio albolineatus) hybrids (cross-breeding between two closely related species interferes with homologous chromosomal pairing during meiosis or induces mitotic arrest [32]).

Zebrafish have been used in many studies to produce germline chimeras [18, 3 for map** results). We calculated the average methylation level for all the genome cytosines in three different genomic contexts. We found that 87% of CpGs were methylated in the sperm samples for all three groups, and the average methylation level for the CHG and CHH contexts ranged between 0.35 and 0.4% (see Additional file 4, cytosine coverage in different contexts). In contrast, the progeny samples exhibited low CpG methylation levels, ranging between 79 and 81%, in which the CHH and CHG contexts accounted for 0.37 to 0.46%. When analysing the genome-wide methylation level, we did not observe a considerable difference between sperm and progeny samples in all three groups (Fig. 2A). Then, we examined methylation density at the whole-genome level. Methylation density showed high conservation among all groups and a trend similar to the methylation level, indicating a close relationship between methylation level and methylation density (Violin plots, Figure S2, Additional file 1). Since CHG and CHH methylation was negligible in all the samples, we focused only on CpG methylation in subsequent analyses.

Fig. 2
figure 2

Global methylation pattern in all three groups. A Violin plots showing methylation levels in all three contexts. The whole genome was split into 10-kb sub-bins, and the methylation level was calculated for each context. The x-axis is labelled with the name of each sample, and the y-axis shows the methylation level. B Circos plots for methylation levels and their differences on each chromosome for the comparison groups. The chromosomes were divided into bins, and each bin’s methylation level was calculated as follows: number of reads with mCs/(number of reads with mCs + number of reads with non-mCs). Description of the circos plot from outside to inside: methylation level for Group 1, colour denotes the methylation level (%); methylation level difference between groups, and the heatmap shows the difference (%); methylation level for Group 2, colour denotes the methylation level (%) and graph legends. The upper panel is for the progeny samples, and the lower panel is for the sperm samples. C The methylation level in the functional regions (promoter, utr5, exon, intron, utr3, CGI, CGI-shore, and repeat elements). After combining the samples with biological replicates, each functional region was divided into 20-kb bins, and the methylation level of each bin was calculated. The upper panel is for the progeny samples, and the lower panel is for the sperm samples. Graph legends contain the compared group names. D Methylation level 2 kb upstream of the TSS, 2 kb downstream of the TES, and the gene body. The upper panel is for the progeny samples, and the lower panel is for the sperm samples. Abbreviations: MOS — sperm derived from the MO chimera; HybS — sperm derived from the hybrid chimera; DoS — sperm derived from the donor; MOH — progeny derived from the MO chimera; HybH — progeny derived from the hybrid chimera; and DoH — progeny derived from the donor, n = 4 biological replicates

Next, the correlation between replicates in each group was calculated using the Pearson correlation coefficient to determine closeness. The whole genome was divided into 2-kb bins, and the number of methylated cytosines (mCpG) was calculated for each bin. Replicates in each group showed a strong correlation, with r2 value > 0.92 (Figure S3, Additional file 1). Next, we checked the correlation between groups of similar sample types. We compared the sperm samples, which showed high homogeneity with r2 value > 0.91, and the result was similar to the progeny samples, in which r2 > 0.9.

Differential methylation analysis suggests no massive epigenetic remodelling in the surrogates

Data from the replicates were combined for the differential methylation analysis between comparison groups at the whole-genome level (Fig. 2B, Circos plot). We observed that the hypermethylation rate in the HybH and MOH groups was slightly higher than that of the DOH group, and the difference in the methylation level was high among all three compared groups. Similarly, the sperm samples from all three groups showed higher differences with a different methylation pattern in all 25 chromosomes. Surprisingly, the landscape of CpG methylation in all genomic functional regions followed a very similar pattern with methylation depletion at the transcription start site (TSS) (Fig. 2C,D). Compared to those in other regions, introns and the repeat elements were hypermethylated and retained a similar shape in the plotted data for all the groups with no substantial difference. Compared to other genomic features, CpG islands (CGIs), which have been reported to predominate the promoter regions in vertebrates and to be involved in regulating gene expression [36, 37], were hypomethylated in all the samples and retained a similar methylation pattern in all studied groups. These results indicate that the sperm and progeny produced by the surrogates did not undergo a high degree of epigenetic remodelling.

DMRs in the compared groups reveal hypermethylation in both surrogates

Since we did not see any differences in the functional regions of the genome, we next investigated the differentially methylated regions (DMRs) in all groups. First, we calculated the number of DMRs in progeny samples across the genome. We found 3788 DMRs between the HybH and DOH, 3814 between the MOH and DOH, and 3766 between the MOH and HybH (Figure S4, Additional file 1). Then, we looked at hypomethylated (hypoDMR) and hypermethylated DMR (hyperDMR) in the compared groups. We annotated the DMRs according to their position in the genome (see Additional file 5 for the number of DMR distributions in all the genomic regions). More DMRs were found in introns and repeat elements than in other functional regions (Fig. 3A). When comparing HybH with DOH, we found more hyperDMRs in all the regions compared to the number of hypoDMRs. Then, we compared MOH and DOH and found almost the same number of both types of DMRs in all regions. However, the scenario was entirely different when we compared MOH with HybH, as we observed a higher number of hypoDMRs in all the regions, indicating higher methylation in HybH. Next, we calculated the methylation level of all the DMRs in the individual groups (Fig. 3B). The median value of the methylation level of the DMRs associated with the HybH was higher than that in the MOH and DOH. When we evaluated the DMR distribution with a violin plot, most DMRs were found to have accumulated with a higher methylation level in MOH and HybH.

Fig. 3
figure 3

DMR methylation level and their distribution in the functional regions. A DMR distribution plot showing the different functional regions in the progeny sample comparison group. The x-axis shows the functional regions, and the y-axis shows the number of hyper-/hypoDMRs in each region. B Violin plot showing the DMR methylation level in the progeny samples in each comparison group. The x-axis shows the comparison group name, and the y-axis shows the methylation level. C DMR distribution in the sperm samples and D DMR methylation level of the sperm samples. The x-axis shows the comparison group name, and the y-axis shows the methylation level

Next, we asked whether the DMRs in sperm samples were similar to those in the progeny groups. To this end, we found 3639 DMRs between the HybS and DOS, 3527 between the MOS and DOS, and 3708 between the MOS and HybS. We annotated the DMRs, assessed their distribution (Fig. 3C), and calculated the methylation levels (Fig. 3D). The DMR distribution in all the sperm groups was similar to that in the progeny groups. The violin plots show that most of the DMRs in HybS and MOS groups exhibited higher methylation levels compared to those in the DOS group (also see the cluster heatmap of the DMRs in Figure S5, Additional file 1). The median value was higher, suggesting that sperm samples from both surrogates carried more hyperDMRs than the donors. These results indicate that both sperm and progeny derived from the surrogates were hypermethylated compared to the methylation level of the donors.

Hypermethylation in promoters regions of protocadherin gene (pcdh) in MOH and HybS

Because most of the DMRs found were accumulated in intronic regions and repeat elements, we focused only on the promoter DMRs to see if any biological function was disrupted in the chimera groups (The list of significant DMPs between the compared group is provided as Additional file 6). Promoter regions are located within 1 kb from the TSS (− 1 to 0). Methylation in the immediate vicinity of a TSS hinders transcription initiation, whereas methylation in the gene body does not and may even stimulate transcription elongation [37]. Hence, we extracted all the differentially methylated promoters (DMPs) and analysed their GO terms; terms with p-value less than 0.05 were considered to be significantly enriched (Figure S6, Additional file 1). To our surprise, in the comparison groups for the progeny samples, more DMPs associated with biological processes were significantly enriched in the MOH compared to HybH and DOH. Biological functions such as homophilic cell adhesion, regulation of odontogenesis, and tooth mineralization were highly enriched in the MOH compared to HybH and DOH. We individually checked all the gene IDs for the promoters associated with this term in the Integrative Genome Viewer (IGV). The identified promoter regions encoded pcdh1gc6, pcdh1g22, pcdh1g30, pcdh1g33, pcdh1g2, and pcdh1g31 (IGV Snapshots for pcdh1g DMPs, Fig. 4A). When visualizing individual replicates from each group of the progeny samples in IGV, the identified DMPs were found to be hypermethylated only in the MOH samples, and the pattern was consistent among the replicates. Next, we wondered whether these marks were present in the sperm samples. We compared the MOS and DOS groups and did not find significantly enriched DMPs related to these genes (Fig. 4B). When checking the methylation pattern in IGV for these DMPs, we see that HybS and DOS have similar methylation level, and the difference only arise in the progeny, where HybH and DOH show low methylation and MOH remains higher, suggesting that these loci in the HybH and DOH were demethylated while those in the MOH remained unchanged. These results indicated that hyperDMPs within the pcdh gene in MOH were transmitted through the germline.

Fig. 4
figure 4

Hypermethylation in the pcdh gene. A IGV snapshots showing multiple representative hyperDMPs in pcdh1g gene, with increased methylation in MOH samples compared to DOH. B IGV view of the same promoter regions showing no hyperDMPs between the MOS and DOS groups. C Predictive hyperDMPs between the HybS and DOS groups in pcdh1a and pcdh2a genes. The black lines below the sample tracks represent predicted DMP regions. DMP regions are also highlighted by red lines above the sample tracks, and RefSeq gene annotations are shown below the tracks. D Relative gene expression levels of pcdh1g2, pcdh1g31, pcdh1g22, pcdh1g33, and itga5 normalized to housekee** gene eef1a1l1 (Individual data values for qPCR are provided in Additional file 7). Data are shown as mean ± SD; One-way ANOVA followed by Tukey’s multiple comparisons for data with Gaussian distribution and Kruskal–Wallis test for data with non-Gaussian distribution (for pcdh1g2 data), p < 0.05, n = 5 biological replicates, ns = not significant

We also found very few functions like cell–cell adhesion and homophilic cell adhesion enrichment for hyperDMPs in HybS, and the gene promoters targeted for these functions are pcdh1g, pcdh2g, and pcdh1a (Fig. 4C). However, the hybrid progenies were unaffected by this hypermethylation since we did not see any function enrichment in the GO analysis. We measured the mRNA expression levels for selected genes, such as pcdh1g2, pcdh1g31, pcdh1g22, and pcdh1g33 by RT‒qPCR (see Fig. 4D, individual data values for qPCR are provided in Additional file 7). However, We did not observe any significant difference in the gene expression level, which suggested that the hyperDMPs in pcdh did not affect the expression of these genes. Next, we analysed the hypoDMPs between MOH and DOH and found that functions such as regulation of angiogenesis, positive regulation of the developmental process, and positive regulation of vasculature development were significantly enriched. Angiogenesis is the process of blood and lymphatic vessel formation during embryonic development [38]. DMPs in genes itga5 and itga6l were found to be hypomethylated. Integrin alpha 5 (itga5) is essential for heart development [39], but RT-qPCR showed no significant difference in the mRNA expression levels for itga5. Next, we examined the hypoDMPs in the comparison groups HybH and DOH for enriched GO terms. We found very few basic enriched molecular functions like nickel cation binding, adenylate kinase activity, nucleotide kinase activity, and phosphotransferase activity. The promoters associated with these terms in the progeny samples were predicted to have transcription factor activity. The genes are si:dkey-14o6.4, si:dkey-8o9.5, and si:dkey-54j5.2 (complete list is provided as Additional file 8). These HypoDMPs were not significant in the sperm samples of any surrogate groups.

Promoter hypermethylation in HybS reveals that MAPK/p53 pathway disruption leads to low-quality sperm production

Next, we investigated the KEGG (Kyoto Encyclopedia of Genes and Genomes) pathway enrichment for DMPs to determine whether any signalling pathway was disrupted. Pathways with a p-value lower than 0.05 in Fisher’s exact test were regarded as significant. When analysing the hypoDMPs between the MOH and DOH samples, we found that metabolic pathways, oxidative phosphorylation, and cardiac muscle contraction were highly enriched functional pathways (Figure S7, Additional file 1). The finding of cardiac muscle contraction enrichment was congruent with GO angiogenesis term enrichment in the MOH group. However, we did not find any significant pathway enrichment in the MOS group. Metabolic pathways are critical mechanisms, and oxidative phosphorylation is an essential function; enriched DMPs were found mainly in the mitochondrial genome (the complete list is provided as Additional file 9). Moreover, mitochondria are maternally inherited by progeny [7). Data are shown as mean ± SD; unpaired t-test for data with Gaussian distribution (mapk12a, mapk8a, map2k2a, and bcl2l1 dataset) and the Mann–Whitney U test for data with non-Gaussian distribution (for Tp53, apaf1, and map4k2 dataset) with adjusted p < 0.05, n = 4 biological replicates, asterisks represent significant difference between the groups, ns = not significant