Split hand/foot malformation (SHFM) is a rare limb abnormality with clefting of the fingers and/or toes. For many individuals, the genetic etiology is unknown. Through whole-exome and targeted sequencing, we detected three novel variants in a gene encoding a transcription factor, PRDM1, that arose de novo in families with SHFM or segregated with the phenotype. PRDM1 is required for limb development; however, its role is not well understood and it is unclear how the PRDM1 variants affect protein function. Using transient and stable overexpression rescue experiments in zebrafish, we show that the variants disrupt the proline/serine-rich and DNA-binding zinc finger domains, resulting in a dominant-negative effect. Through gene expression assays, RNA sequencing, and CUT&RUN in isolated pectoral fin cells, we demonstrate that Prdm1a directly binds to and regulates genes required for fin induction, outgrowth and anterior/posterior patterning, such as fgfr1a, dlx5a, dlx6a and smo. Taken together, these results improve our understanding of the role of PRDM1 in the limb gene regulatory network and identified novel PRDM1 variants that link to SHFM in humans.

Vertebrate limb development is controlled by a complex gene regulatory network (GRN) governed by signaling pathways, transcription factors and epigenetic modifiers. Limb growth begins at the lateral plate mesoderm, where mesenchyme precursors form a small bud surrounded by an ectodermal layer. Retinoic acid and Wnt signaling initiate limb induction (Grandel et al., 2002; Ng et al., 2002), and outgrowth is driven by the apical ectodermal ridge (AER), where transcription factors TBX5 and TP63 induce expression of fibroblast growth factor 10 (Fgf10) in the mesenchyme and Fgf8 in the outer ectoderm (Agarwal et al., 2003; Bakkers et al., 2002; Ng et al., 2002). This establishes a complex epithelial-mesenchymal feedback loop that then activates proliferation and differentiation of mesenchymal cells for limb growth (proximal/distal axis) (Ohuchi et al., 1997). Anterior/posterior patterning, or establishment of digits 1-5, is regulated by sonic hedgehog (Shh) signaling in the zone of polarizing activity (ZPA) (Riddle et al., 1993; Saunders and Gasseling, 1968). Each gene and pathway are interconnected, and dysregulation at any point, particularly in the AER, can cause abnormal limb growth (Kantaputra and Carlson, 2019).

Misregulation of the limb GRN can lead to congenital limb defects, which affect one in 2000 newborns (Wilcox et al., 2015). Split hand/foot malformation (SHFM) is a limb abnormality resulting in missing, hypoplastic and/or fused digits. SHFM occurs in one in 18,000 live births and there are eight known forms of the disease due to pathogenic variants in WNT10B (MIM #225300), TP63 (MIM #605289), DLX5 (MIM #183600), ZAK (or MAP3K20; MIM #616890) or EPS15L1 (MIM *616826), or chromosomal rearrangements in chromosomes 2 (MIM %606708), 10 (MIM #246560) or X (MIM %313350) (reviewed in Umair and Hayat, 2020). However, in 50% of cases, the genetic etiology is unknown (Sowinska-Seidler et al., 2014). Deletions and translocations at 6q21 have also been associated with SHFM, although no candidate gene has been isolated prior to now (Braverman et al., 1993; Correa-Cerro et al., 1996; Duran-Gonzalez et al., 2007; Gurrieri et al., 1995; Hopkin et al., 1997; Pandya et al., 1995; Tsukahara et al., 1997; Viljoen and Smart, 1993). Here, we report three families with SHFM of unknown genetic etiology, and using whole-exome sequencing (WES) and targeted sequencing, we identified three different variants of unknown significance in a gene encoding a transcription factor, PRDM1, located at 6q21.

PRDM1, also known as BLIMP1 (MIM *603423), is required for limb development, although its role not well understood (Ha and Riddle, 2003; Lee and Roy, 2006; Mercader et al., 2006; Robertson et al., 2007; Vincent et al., 2005; Wilm and Solnica-Krezel, 2005). The protein has an N-terminal SET domain, followed by a proline/serine-rich domain, and five zinc fingers. In various contexts, PRDM1 can bind to DNA through its zinc finger domains and activate or repress gene expression (reviewed in Bikoff et al., 2009; Powell et al., 2013). SET domains are often associated with histone methyltransferase activity (Cheng et al., 2005; Martin and Zhang, 2005), but this has not been observed for PRDM1 in vivo (Hohenauer and Moore, 2012). Rather, it can indirectly alter transcription by forming complexes with chromatin-modifying proteins, such as the histone demethylase Kdm4a (Prajapati et al., 2019), histone methyltransferases Prmt5 (Ancelin et al., 2006) and G9a (or EHMT2) (Gyory et al., 2004), and histone deacetylases HDAC1/2 (Yu et al., 2000) at both the proline/serine and zinc finger domains (reviewed in Bikoff et al., 2009).

Studies in mice and zebrafish indicate that PRDM1 is important for limb and pectoral fin formation. Homozygous Blimp1GFP/GFP mice, which have a STOP-IRESgfp-pgk neo cassette inserted after exon 6, eliminating expression of the PRDM1 zinc fingers, results in a loss of digit 5 and a shortened ulna (Kallies et al., 2004; Robertson et al., 2007). Transheterozygous Blimp1GFP/− mice lack posterior digits 4 and 5 and an ulna, whereas conditional knockouts of Prdm1 in the embryo (Sox2:Cre) causes loss of posterior digits 3-5 and the ulna, owing to disruption of sonic hedgehog signaling and dysregulation in the ZPA (Robertson et al., 2007). This suggests that a graded loss of Prdm1 in mice results in gradually more severe limb phenotypes. Zebrafish embryos injected with prdm1a morpholinos (‘morphants’) for targeted knockdown fail to develop a pectoral fin, which is homologous to mammalian forelimbs, whereas a hypomorphic allele, prdm1atp39/tp39, presents with mild phenotypes, namely, shortening of the scapulocoracoid and variable truncation of the fin overall (Lee and Roy, 2006; Mercader et al., 2006). These studies show that Prdm1a is downstream of tbx5a and upstream of fgf10a during fin induction. It is also required for shha activity in the ZPA (Lee and Roy, 2006; Mercader et al., 2006). However, it is unclear whether this is by direct transcriptional regulation or by recruitment of epigenetic modifiers. Although PRDM1/Prdm1a has been shown to be important in limb and pectoral fin development, how it functions molecularly is poorly understood.

We sought to better understand the mechanistic role of PRDM1 in limb development and SHFM. We identified novel PRDM1 variants in families with SHFM and show through transient and stable overexpression assays in zebrafish that these variants act in a dominant-negative fashion due to disruption of the proline/serine and DNA-binding zinc finger domains. We used RNA sequencing (RNA-seq) and Cleavage Under Targets and Release Using Nuclease (CUT&RUN; Meers et al., 2019; Skene and Henikoff, 2017; Ye et al., 2021) in isolated pectoral fin cells to show that Prdm1a directly binds to regulatory sequences of fgfr1a, dlx5a, dlx6a and smo, and regulates their expression in the fin. These data show that Prdm1a is involved in fin induction, outgrowth and anterior/posterior patterning and requires its proline/serine and zinc finger domains to accomplish these morphogenic processes. Taken together, these results improve our understanding of the role of PRDM1 in the limb GRN, introduce novel SHFM PRDM1 alleles, and help us better predict the pathogenicity of PRDM1 variants in humans.

WES in individuals with SHFM reveals novel PRDM1 variants of unknown significance

SHFM is a congenital limb disorder in which individuals exhibit missing, shortening or fusions of the fingers and toes. Phenotypes vary due to incomplete penetrance, and for 50% of individuals, the genetic etiology is unknown (Sowinska-Seidler et al., 2014). We performed WES on a multi-generational family with SHFM that was negative for TP63 variants and whose single-nucleotide polymorphism microarray appeared normal. Four individuals were heterozygous for a mutation in PRDM1: PRDM1c.712_713insT (p.C239Lfs*32), which introduces a single base-pair insertion causing a frameshift and premature stop codon after the SET domain as well as predicted truncation of the protein (Fig. 1A,B). The probability of being loss-of-function intolerant (pLI score) for PRDM1 on gnomAD v.2.1.1 is 0.96 (Karczewski et al., 2020), suggesting that the gene is intolerant to loss-of-function variants. This variant is not observed in gnomAD v2.1.1, is predicted to be pathogenic by MutationTaster (https://www.mutationtaster.org/), and is the only gene from the WES results known to be involved in limb development (Fig. 1A; Table S1) (Ha and Riddle, 2003; Lee and Roy, 2006; Mercader et al., 2006; Robertson et al., 2007; Vincent et al., 2005; Wilm and Solnica-Krezel, 2005). As expected, SHFM in this family is variable (Fig. 1C). Individuals II:3 and III:3 have missing digits and clinodactyly but apparently normal feet. Individual III:3 has a clefted hand and missing toe. Individual III:4 has a mild phenotype of minor brachydactyly (shortened digits). Individuals IV:1 and IV:2 are monozygotic twins, although only IV:2 has SHFM (missing digits and clefting) (Fig. 1C). This may be due to environmental differences while in utero, epigenetic differences and/or the presence of genetic modifiers (Castillo-Fernandez et al., 2014; Gordon et al., 2012). Although the phenotypes within the family are variable, they were generally expected. In mice, removing PRDM1, or even just the zinc finger domain of PRDM1, results in missing posterior digits, as seen in these SHFM individuals (Robertson et al., 2007). Variants in PRDM1 may cause SHFM in an autosomal-dominant pattern with variable penetrance and expressivity.

Fig. 1.

PRDM1 variants of unknown significance identified in families with split hand/foot malformation (SHFM). (A) Table showing PRDM1 variants and predictions of pathogenicity based on various bioinformatics tools. CADD, Combined Annotation-Dependent Depletion; FS, frameshift; MAF, minor allele frequency; MutTat, MutationTaster; NA, not applicable; PP2_HD, Polymorphism Phenotyping v2 HumDiv; SIFT, Sorting Intolerant From Tolerant. (B) Schematic of PRDM1 structure and location of variants identified in individuals with SHFM. (C) Pedigree for family with PRDM1 variant #1, c.712_713insT (p.C239Lfs*32). The symbols representing affected individuals are shaded. Standard pedigree symbols are used. The variant is inherited in an autosomal-dominant manner with incomplete penetrance and variable expressivity. Photographs of the limbs of the individuals in the family are also shown. (D) Pedigree for family with PRDM1 variant #2, c.1571C>G (p.T524R). (E) Pedigree for family with PRDM1 variant #3, c.2455A>G (p.T819A). Variants #2 and #3 are de novo.

Fig. 1.

PRDM1 variants of unknown significance identified in families with split hand/foot malformation (SHFM). (A) Table showing PRDM1 variants and predictions of pathogenicity based on various bioinformatics tools. CADD, Combined Annotation-Dependent Depletion; FS, frameshift; MAF, minor allele frequency; MutTat, MutationTaster; NA, not applicable; PP2_HD, Polymorphism Phenotyping v2 HumDiv; SIFT, Sorting Intolerant From Tolerant. (B) Schematic of PRDM1 structure and location of variants identified in individuals with SHFM. (C) Pedigree for family with PRDM1 variant #1, c.712_713insT (p.C239Lfs*32). The symbols representing affected individuals are shaded. Standard pedigree symbols are used. The variant is inherited in an autosomal-dominant manner with incomplete penetrance and variable expressivity. Photographs of the limbs of the individuals in the family are also shown. (D) Pedigree for family with PRDM1 variant #2, c.1571C>G (p.T524R). (E) Pedigree for family with PRDM1 variant #3, c.2455A>G (p.T819A). Variants #2 and #3 are de novo.

We then screened an additional 75 unrelated people with SHFM and performed targeted sequencing for PRDM1. One individual has ectrodactyly ectodermal dysplasia (EEC) syndrome (MIM 129810) with bilateral 3/4-digit syndactyly and a high arch palate but no clefting. Testing for TP63 variants was normal, but there was a missense variant in PRDM1: PRDM1c.1571C>G (p.T524R) (Fig. 1A,D). Another SHFM individual has bilateral tibial deficiency with shortening and clubfoot. This individual was normal for TP63, SNX3 (MIM %601349) and NR2E1 (Kumar et al., 2007) variants but presented with a missense variant, PRDM1c.2455A>G (p.T819A) (Fig. 1A,E). Both individuals were heterozygous, and the variants were absent from both sets of parents based on targeted sequencing, suggesting a de novo mutation (Fig. 1D,E). Both variants have a minor allele frequency of ≤6.6×10−4 (gnomAD v.2.1.1) and are predicted to be pathogenic by at least two dbNSFP tools (https://sites.google.com/site/jpopgen/dbNSFP) (Fig. 1A) (Liu et al., 2020). Additionally, they flank the zinc finger domain of PRDM1 and may result in a loss of phosphorylation at these sites or affect protein folding and its ability to bind DNA (Fig. 1B) (Keller and Maniatis, 1992). Interestingly, these two missense variants did not result in digit loss, suggesting a milder effect on the zinc finger domain than that of the first allele. Taken together, our data suggest that variants in PRDM1 underlie SHFM.

Loss of Prdm1a causes pectoral fin defects

PRDM1 is required for vertebrate limb development (Ha and Riddle, 2003; Lee and Roy, 2006; Mercader et al., 2006; Robertson et al., 2007; Vincent et al., 2005; Wilm and Solnica-Krezel, 2005). The zebrafish pectoral fin is homologous to mammalian forelimbs and its early structures consist of a cleithrum, scapulocoracoid/postcoracoid, endoskeletal disk and fin fold, which are all derived from mesenchymal cells (Fig. 2A) (Grandel and Schulte-Merker, 1998). In zebrafish, prdm1a is first expressed in the pectoral fin at 18 h post fertilization (hpf). It is highly expressed in the fin mesenchyme, pharyngeal arches and neurons at 24 hpf (Fig. 2B). At 48 hpf, expression continues in the pharyngeal arches and neurons, and prdm1a is present in the apical fold (AF) (Wilm and Solnica-Krezel, 2005) (Fig. 2C), which is analogous to the AER (reviewed in Yano and Tamura, 2013). Previous studies have shown that knockdown of prdm1a using anti-sense morpholinos completely disrupts pectoral fin growth (Mercader et al., 2006). Hypomorph prdm1atp39/tp39 mutants, which have a missense mutation (p.H564R) in the second zinc finger (Fig. 2D), present with mild phenotypes, namely a shortening of the scapulocoracoid and variable truncation of the fin, which is incompletely penetrant, occurring in only 30% of mutants (Baxendale et al., 2004; Lee and Roy, 2006; Roy et al., 2001). Predicted null prdm1am805/m805 mutants, hereafter referred to as prdm1a−/−, have a mutation resulting in a premature stop codon in the SET domain (p.W154*) (Fig. 2D) (Artinger et al., 1999; Hernandez-Lagunas et al., 2005). To evaluate fin phenotypes in prdm1a−/− mutants, embryos were stained with Alcian Blue to assess cartilage development at 4 days post fertilization (dpf). There were no significant differences between wild-type (WT) and prdm1a+/− heterozygotes (Fig. S1). Heterozygotes were included in all WT measurements. prdm1a−/− mutants presented with pronounced pectoral fin defects (Fig. 2E-J). There was a significant decrease in the average length of the cleithrum (∼20% decrease, P=0.0173), endoskeletal disk (8.7%, P=0.0816) and fin fold (10.6%, P=0.0374) (Fig. 2G-J). We performed immunostaining for phosphorylated histone H3 and cleaved caspase 3 at 48 hpf in the pectoral fins to mark cell proliferation and cell death, respectively. prdm1a−/− mutants had a slight decrease in cell proliferation and an increase in cell death, although the results were not statistically significant (Fig. S2A-D). These results suggest that Prdm1a may be involved in cell differentiation. In addition, they indicate that the zinc finger domain is important for the function of Prdm1a in pectoral fin development and is specifically required for differentiation of the skeletal elements.

Fig. 2.

prdm1a−/− zebrafish mutants have hypoplastic pectoral fins. (A) Cartoon of pectoral fin bud at 4 dpf. (B,C) Lateral view of whole-mounted embryos after hybridization chain reaction (HCR) for prdm1a at (B) 24 (n=3) and (C) 48 hpf (n=28). Anterior is to the left. Images are maximum projections of the whole embryo. Arrowheads point to the pectoral fin bud. Scale bars: 100 µm. (D) Schematic of Prdm1a protein. The prdm1am805/m805 allele causes a premature stop codon in the SET domain and is a presumed null mutation (p.W154*). The hypomorphic prdm1atp39/tp39 allele is a missense mutation in the second zinc finger (p.H564R). (E,F) Representative images of Alcian Blue-stained pectoral fins for (E) WT/heterozygous (n=18) and (F) prdm1a−/− mutants (n=9) at 4 dpf. Scale bars: 50 µm. (G-J) The average lengths of the (G) cleithrum, (H) endoskeletal disk and (I) fin fold and (J) the average area of the scapulocoracoid/postcoracoid were measured. Each dot represents one independent biological replicate. Averages were compared with an unpaired, two-tailed independent Student's t-test. prdm1a−/− mutants had a shorter cleithrum (P=0.0173), endoskeletal disk (P=0.0816) and fin fold (P=0.0374). Error bars represent the mean±s.d. The representative images in E and F are also shown in Fig. S1, where the WT and heterozygous fins are analyzed separately. Ant, anterior; cl, cleithrum; D, distal; e, eye; ed, endoskeletal disk; ff, fin fold; Het, heterozygous; hpf, hours post fertilization; MT, prdm1a−/− mutant; n, neurons; P, proximal; pa, pharyngeal arches; pc, postcoracoid; pf, pectoral fin; Post, posterior; sc, scapulocoracoid; WT, wild-type; y, yolk.

Fig. 2.

prdm1a−/− zebrafish mutants have hypoplastic pectoral fins. (A) Cartoon of pectoral fin bud at 4 dpf. (B,C) Lateral view of whole-mounted embryos after hybridization chain reaction (HCR) for prdm1a at (B) 24 (n=3) and (C) 48 hpf (n=28). Anterior is to the left. Images are maximum projections of the whole embryo. Arrowheads point to the pectoral fin bud. Scale bars: 100 µm. (D) Schematic of Prdm1a protein. The prdm1am805/m805 allele causes a premature stop codon in the SET domain and is a presumed null mutation (p.W154*). The hypomorphic prdm1atp39/tp39 allele is a missense mutation in the second zinc finger (p.H564R). (E,F) Representative images of Alcian Blue-stained pectoral fins for (E) WT/heterozygous (n=18) and (F) prdm1a−/− mutants (n=9) at 4 dpf. Scale bars: 50 µm. (G-J) The average lengths of the (G) cleithrum, (H) endoskeletal disk and (I) fin fold and (J) the average area of the scapulocoracoid/postcoracoid were measured. Each dot represents one independent biological replicate. Averages were compared with an unpaired, two-tailed independent Student's t-test. prdm1a−/− mutants had a shorter cleithrum (P=0.0173), endoskeletal disk (P=0.0816) and fin fold (P=0.0374). Error bars represent the mean±s.d. The representative images in E and F are also shown in Fig. S1, where the WT and heterozygous fins are analyzed separately. Ant, anterior; cl, cleithrum; D, distal; e, eye; ed, endoskeletal disk; ff, fin fold; Het, heterozygous; hpf, hours post fertilization; MT, prdm1a−/− mutant; n, neurons; P, proximal; pa, pharyngeal arches; pc, postcoracoid; pf, pectoral fin; Post, posterior; sc, scapulocoracoid; WT, wild-type; y, yolk.

SHFM human PRDM1 variants have a dominant negative effect

To test whether the SHFM human PRDM1 (hereafter hPRDM1) variants are functional, we designed an in vivo pectoral fin rescue experiment in which hPRDM1 variants were overexpressed. We overexpressed either WT hPRDM1 mRNA or each of the three SHFM variants in intercrossed prdm1a+/− zebrafish embryos (Fig. 3A-F). Embryos were staged throughout the first four days of development to ensure that there was no developmental delay or unassociated pathologies due to the mRNA overexpression. In prdm1a−/− mutants, injection of WT hPRDM1 partially rescued the pectoral fin defects and the fins more closely resembled those of uninjected WT (Fig. 3A-C), particularly the length of the cleithrum (12% increase), endoskeletal disk (5.5%), and fin fold (6%) (Fig. 3G-J). Inability of the WT allele to fully rescue the pectoral fin was likely due to the transience of the assay and rapid degradation of the mRNA. In contrast, overexpression of the three SHFM variants failed to rescue the elements of the pectoral fin (Fig. 3D-F). Indeed, injection of the hPRDM1 variant p.T819A in mutants further exacerbated hypoplasia of the endoskeletal disk (P=0.0138) (Fig. 3F,H). These data suggest that the SHFM variants are pathogenic compared to the WT allele. Overexpression of the SHFM variants hPRDM1 p.C239Lfs*32 and p.T819A led to a significant decrease in endoskeletal disk length in WT embryos (P=0.0135 and P=0.0280, respectively), suggesting a dominant-negative effect of the alleles on pectoral fin development, although we cannot rule out a hypermorphic or neomorphic function (Fig. S3). Given the location of the alleles, we predict that the zinc finger domain is important for the function of PRDM1 in limb development.

Fig. 3.

Transient overexpression of SHFM hPRDM1 variants fails to rescue the pectoral fin in prdm1a−/− mutants.prdm1a+/− heterozygous fish were intercrossed and injected with the hPRDM1 WT and SHFM variant mRNAs at the single-cell stage. Injected larvae were collected at 4 dpf for Alcian Blue staining. (A-F) Representative images of Alcian Blue-stained pectoral fins at 4 dpf. (A) Uninjected WT/heterozygous (n=18). (B) Uninjected prdm1a−/− mutant (n=9). (C-F) prdm1a−/− mutants were injected with (C) WT hPRDM1 (n=10), (D) hPRDM1(p.C239Lfs*32) (n=10), (E) hPRDM1(p.T524R) (n=8) or (F) hPRDM1(p.T819A) mRNA (n=7). The representative uninjected control images in A and B are also used in Fig. S3, which shows the effect of overexpression of hPRDM1 variants in the WT background as part of the same experiment. Scale bars: 50 µm. (G-I) Measurements for the lengths of the (G) cleithrum, (H) endoskeletal disk and (I) fin fold and (J) the area of the scapulocoracoid and postcoracoid were averaged and compared using a one-way ANOVA, followed by a Tukey post-hoc test relative to uninjected prdm1a−/− mutants. Each dot represents one independent biological replicate. P-values are shown in the figure. Injection of WT hPRDM1 partially rescued the cleithrum, endoskeletal disk and fin fold of prdm1a−/− mutants. However, overexpression of the three SHFM variants failed to rescue the pectoral fin. Error bars represent the mean±s.d. cl, cleithrum; D, distal; ed, endoskeletal disk; ff, fin fold; Het, heterozygous; hpf, hours post fertilization; MT, prdm1a−/− mutant; P, proximal; pc, postcoracoid; sc, scapulocoracoid; WT, wildtype.

Fig. 3.

Transient overexpression of SHFM hPRDM1 variants fails to rescue the pectoral fin in prdm1a−/− mutants.prdm1a+/− heterozygous fish were intercrossed and injected with the hPRDM1 WT and SHFM variant mRNAs at the single-cell stage. Injected larvae were collected at 4 dpf for Alcian Blue staining. (A-F) Representative images of Alcian Blue-stained pectoral fins at 4 dpf. (A) Uninjected WT/heterozygous (n=18). (B) Uninjected prdm1a−/− mutant (n=9). (C-F) prdm1a−/− mutants were injected with (C) WT hPRDM1 (n=10), (D) hPRDM1(p.C239Lfs*32) (n=10), (E) hPRDM1(p.T524R) (n=8) or (F) hPRDM1(p.T819A) mRNA (n=7). The representative uninjected control images in A and B are also used in Fig. S3, which shows the effect of overexpression of hPRDM1 variants in the WT background as part of the same experiment. Scale bars: 50 µm. (G-I) Measurements for the lengths of the (G) cleithrum, (H) endoskeletal disk and (I) fin fold and (J) the area of the scapulocoracoid and postcoracoid were averaged and compared using a one-way ANOVA, followed by a Tukey post-hoc test relative to uninjected prdm1a−/− mutants. Each dot represents one independent biological replicate. P-values are shown in the figure. Injection of WT hPRDM1 partially rescued the cleithrum, endoskeletal disk and fin fold of prdm1a−/− mutants. However, overexpression of the three SHFM variants failed to rescue the pectoral fin. Error bars represent the mean±s.d. cl, cleithrum; D, distal; ed, endoskeletal disk; ff, fin fold; Het, heterozygous; hpf, hours post fertilization; MT, prdm1a−/− mutant; P, proximal; pc, postcoracoid; sc, scapulocoracoid; WT, wildtype.

Prdm1a proline/serine-rich and DNA-binding zinc finger domains are required to regulate pectoral fin development

PRDM1 has an N-terminal SET domain, followed by a proline/serine-rich domain, and five zinc fingers. To determine the functionally active domain of Prdm1a during pectoral fin development, we overexpressed modified versions of Prdm1a in null mutants using a stable, conditional Gal4/UAS system (Fig. 4A,B). Transgenic fish expressing Gal4 under a heat-shock promoter, Tg(hsp70l:gal4)co1025Tg, were generated and crossed to prdm1a+/− to create Tg(hsp70I:gal4);prdm1a+/− fish. Using site-directed mutagenesis, we deleted each of the three functional domains of Prdm1a (Fig. 4A; Table S2). These deletions were modeled after previous in vitro studies (Gyory et al., 2004; Ren et al., 1999; Su et al., 2009; Yu et al., 2000). The modified genes were tagged with a self-cleaving 2a-EGFP reporter and placed under the control of a 4Xnr UAS enhancer. At the single-cell-stage, we injected the 4XnrUAS-modified prdm1a-2a-EGFP constructs into Tg(hsp70l:gal4);prdm1a+/− intercrossed embryos along with Tol2 transposase mRNA. During normal development, prdm1a is first expressed in the pectoral fin at 18 hpf. Therefore, we heat shocked the embryos at 6 hpf (shield stage), giving the embryos time to transcribe, translate and activate the Gal4 protein. The Gal4 protein would then bind to the UAS and activate transcription of the modified prdm1a construct. At 24 hpf, we screened embryos for mosaic EGFP expression (Fig. 4C,D), and then stained them with Alcian Blue at 4 dpf to assess the level of rescue to the pectoral fin compared to uninjected controls (Fig. 4E-L; Fig. S4). Of note, mosaic EGFP expression in injected and heat-shocked embryos was highly variable and may have played a role in the ability of the construct to rescue the pectoral fin. prdm1a−/− mutants injected with the positive control, the construct expressing full-length Prdm1a, exhibited a rescue, particularly in the endoskeletal disk (25.6% increase, P=0.0654), compared to uninjected controls (Fig. 4F,G,N). Deletion of the SET domain (Prdm1aΔSET) also partially rescued the area of the scapulocoracoid/postcoracoid (38.0% increase, P=0.0446) (Fig. 4H,P), cleithrum (25.8%, P=0.1252) (Fig. 4M) and endoskeletal disk (22.9%) (Fig. 4N), suggesting that this domain is not important for pectoral fin development. However, when either the proline/serine (Prdm1aΔP/S) or zinc finger (Prdm1aΔZnF) domain was deleted, the constructs failed to rescue the pectoral fin (Fig. 4I,J,M-P). Furthermore, when the two domains were deleted together (Prdm1aΔP/S&ZnF), we failed to see a rescue in any of the structures (Fig. 4K,M-P). The proline/serine and zinc finger domains are important for recruiting epigenetic modifiers and binding DNA, respectively. Injection of a negative control, a construct expressing EGFP, also failed to rescue any cartilage structures in the pectoral fin of prdm1a−/− mutants (Fig. 4L-P). Our data suggest that Prdm1a requires both its proline/serine and zinc finger domains to properly regulate fin development. These data highlight the importance of the proline/serine and zinc finger domains, which were disrupted in the SHFM families in this study.

Fig. 4.

Overexpression of Prdm1a using a global heat-shock Gal4/UAS system shows that proline/serine-rich and zinc finger domains are required for pectoral fin function. (A) Schematic of 4XnrUAS-modified prdm1a-2a-EGFP constructs that were injected into Tg(hsp70l:gal4FF);prdm1a+/− intercrosses. Results for the ability to rescue the pectoral fin are shown. (B) Experimental design for heat-shock Gal4/UAS rescue experiments. Following injection with the UAS construct, embryos at 6 hpf (shield stage) were heat shocked, leading to activation of Gal4, expression of the 4XnrUAS-modified prdm1a-2a-EGFP construct, and cleavage of the 2a viral peptide from EGFP. Embryos were screened for mosaic EGFP expression at 24 hpf. (C,D) Representative images of 24 hpf embryos injected with the 4XnrUAS-modified prdm1a-2a-EGFP construct at the single-cell stage. (C) No heat shock (control). (D) Mosaic EGFP expression in embryos that were injected and heat shocked. The dotted box marks the pectoral fin. Scale bars: 200 µm. (E-L) Representative images of Alcian Blue-stained pectoral fins at 4 dpf are shown. (E) Uninjected WT (n=36). (F) Uninjected prdm1a−/− mutants (n=11). (G-L) Mutants were injected with constructs containing (G) full-length Prdm1a (n=9), (H) Prdm1aΔSET (n=7), (I) Prdm1aΔP/S (n=13), (J) Prdm1aΔZnF (n=7), (K) Prdm1aΔP/S&ZnF (n=13) and (L) an EGFP negative control (n=16). Scale bars: 50 µm. The representative uninjected control images in E and F are also used in Fig. S4, which shows the effect of overexpression of modified Prdm1a in the WT background as part of the same experiment. (M-P) Measurements were taken for the length of the (M) cleithrum, (N) endoskeletal disk and (O) fin fold and (P) the area of the scapulocoracoid and postcoracoid. Each dot represents one independent biological replicate. Measurements for each individual were averaged and compared using a one-way ANOVA, followed by a Tukey's post-hoc test relative to uninjected, heat-shocked prdm1a−/− mutants. prdm1a−/− mutants injected with full-length Prdm1a exhibited a rescue in the endoskeletal disk (P=0.0654). Prdm1aΔSET also partially rescued the area of the scapulocoracoid/postcoracoid (P=0.0446). However, Prdm1aΔP/S, Prdm1aΔZnF and Prdm1aΔP/S&ZnF failed to rescue prdm1a−/− mutants. Error bars represent the mean±s.d. Δ, deleted; Ant, anterior; cl, cleithrum; D, distal; dpf, days post fertilization; e, eye; ed, endoskeletal disk; ff, fin fold; hpf, hours post fertilization; MT, prdm1a−/− mutant; P, proximal; pc, postcoracoid; pf, pectoral fin; Post, posterior; sc, scapulocoracoid; WT, wildtype; y, yolk.

Fig. 4.

Overexpression of Prdm1a using a global heat-shock Gal4/UAS system shows that proline/serine-rich and zinc finger domains are required for pectoral fin function. (A) Schematic of 4XnrUAS-modified prdm1a-2a-EGFP constructs that were injected into Tg(hsp70l:gal4FF);prdm1a+/− intercrosses. Results for the ability to rescue the pectoral fin are shown. (B) Experimental design for heat-shock Gal4/UAS rescue experiments. Following injection with the UAS construct, embryos at 6 hpf (shield stage) were heat shocked, leading to activation of Gal4, expression of the 4XnrUAS-modified prdm1a-2a-EGFP construct, and cleavage of the 2a viral peptide from EGFP. Embryos were screened for mosaic EGFP expression at 24 hpf. (C,D) Representative images of 24 hpf embryos injected with the 4XnrUAS-modified prdm1a-2a-EGFP construct at the single-cell stage. (C) No heat shock (control). (D) Mosaic EGFP expression in embryos that were injected and heat shocked. The dotted box marks the pectoral fin. Scale bars: 200 µm. (E-L) Representative images of Alcian Blue-stained pectoral fins at 4 dpf are shown. (E) Uninjected WT (n=36). (F) Uninjected prdm1a−/− mutants (n=11). (G-L) Mutants were injected with constructs containing (G) full-length Prdm1a (n=9), (H) Prdm1aΔSET (n=7), (I) Prdm1aΔP/S (n=13), (J) Prdm1aΔZnF (n=7), (K) Prdm1aΔP/S&ZnF (n=13) and (L) an EGFP negative control (n=16). Scale bars: 50 µm. The representative uninjected control images in E and F are also used in Fig. S4, which shows the effect of overexpression of modified Prdm1a in the WT background as part of the same experiment. (M-P) Measurements were taken for the length of the (M) cleithrum, (N) endoskeletal disk and (O) fin fold and (P) the area of the scapulocoracoid and postcoracoid. Each dot represents one independent biological replicate. Measurements for each individual were averaged and compared using a one-way ANOVA, followed by a Tukey's post-hoc test relative to uninjected, heat-shocked prdm1a−/− mutants. prdm1a−/− mutants injected with full-length Prdm1a exhibited a rescue in the endoskeletal disk (P=0.0654). Prdm1aΔSET also partially rescued the area of the scapulocoracoid/postcoracoid (P=0.0446). However, Prdm1aΔP/S, Prdm1aΔZnF and Prdm1aΔP/S&ZnF failed to rescue prdm1a−/− mutants. Error bars represent the mean±s.d. Δ, deleted; Ant, anterior; cl, cleithrum; D, distal; dpf, days post fertilization; e, eye; ed, endoskeletal disk; ff, fin fold; hpf, hours post fertilization; MT, prdm1a−/− mutant; P, proximal; pc, postcoracoid; pf, pectoral fin; Post, posterior; sc, scapulocoracoid; WT, wildtype; y, yolk.

Prdm1a controls Fgf signaling in the fin mesenchyme and maintenance of outgrowth and patterning genes

To better understand the disease state of the SHFM individuals and to determine the molecular mechanism by which Prdm1a regulates pectoral fin development, transgenic fish expressing EGFP under a mouse Prx1 (or Prrx1) enhancer, Tg(Mmu:Prx1-EGFP)co1026Tg, were generated and crossed to prdm1a+/− to create Tg(Mmu:Prx1-EGFP);prdm1a+/− heterozygous fish. At 48 hpf, this transgene labels the pectoral fin, pharyngeal arches and dorsal part of the head with EGFP (Fig. 5A,A′) (Hernández-Vega and Minguillón, 2011; Yano and Tamura, 2013). Tg(Mmu:Prx1-EGFP);prdm1a+/− fish were intercrossed and, at 48 hpf, WT and prdm1a−/− embryos were dissected to remove the head and pharyngeal arches. EGFP-positive pectoral fin cells were isolated using fluorescence-activated cell sorting (FACS) before they were subjected to bulk RNA-seq on the Illumina NovaSEQ 6000 system.

Fig. 5.

Loss of Prdm1a leads to downregulation of important limb development genes in the pectoral fin. RNA-seq was performed on isolated pectoral fin cells from about 250 WT and prdm1a−/− embryos at 48 hpf. (A) Lateral and (A′) dorsal view of EGFP-positive pectoral fins from the Tg(Mmu:Prx1-EGFP) zebrafish line at 48 hpf before FACS. Dashed lines indicate where the embryos were dissected prior to FACS. Scale bars: 200 µm. (B) Heat map of top 250 differentially expressed genes (Padj) between WT and prdm1a−/− embryos. (C) Volcano plot showing spread of differentially expressed genes in pectoral fins of prdm1a−/− compared to WT embryos. Light blue dots are significant, differentially expressed genes [−log10(P-value)≥1.15]. Purple dots are downregulated, selected genes of interest, whereas yellow dots are upregulated genes. (D) Downregulated genes in prdm1a−/− embryos were subjected to GO (Panther) pathway enrichment analysis. Yellow arrowheads highlight pathways of interest. A, anterior; ba, branchial arches; DEG, differentially expressed genes; e, eye; P, posterior; pf, pectoral fin; y, yolk.

Fig. 5.

Loss of Prdm1a leads to downregulation of important limb development genes in the pectoral fin. RNA-seq was performed on isolated pectoral fin cells from about 250 WT and prdm1a−/− embryos at 48 hpf. (A) Lateral and (A′) dorsal view of EGFP-positive pectoral fins from the Tg(Mmu:Prx1-EGFP) zebrafish line at 48 hpf before FACS. Dashed lines indicate where the embryos were dissected prior to FACS. Scale bars: 200 µm. (B) Heat map of top 250 differentially expressed genes (Padj) between WT and prdm1a−/− embryos. (C) Volcano plot showing spread of differentially expressed genes in pectoral fins of prdm1a−/− compared to WT embryos. Light blue dots are significant, differentially expressed genes [−log10(P-value)≥1.15]. Purple dots are downregulated, selected genes of interest, whereas yellow dots are upregulated genes. (D) Downregulated genes in prdm1a−/− embryos were subjected to GO (Panther) pathway enrichment analysis. Yellow arrowheads highlight pathways of interest. A, anterior; ba, branchial arches; DEG, differentially expressed genes; e, eye; P, posterior; pf, pectoral fin; y, yolk.

Our RNA-seq analysis revealed a total of 1476 differentially expressed genes between WT and prdm1a−/− mutants specifically in the pectoral fin [−log10(P-value)≥1.2]. Of these, 768 were upregulated, whereas 708 were downregulated (Fig. 5B,C). The most significant downregulated gene was emilin3a, a glycoprotein within the extracellular matrix belonging to the emilin/multimerin family, which, to date, has been shown to be expressed in the notochord, pharyngeal arches and developing craniofacial skeleton of zebrafish (Corallo et al., 2013; Milanetto et al., 2007). The Tg(Mmu:Prx1-EGFP) transgenic line used for the RNA-seq also expresses EGFP in the pharyngeal arches and dorsal part of the head at 48 hpf (Hernández-Vega and Minguillón, 2011; Yano and Tamura, 2013). Although we dissected and removed these regions prior to FACS, there may have been some residual arch and dorsal head tissue in our samples. Key genes known to be involved in pectoral fin development, including members of the hoxa and hoxd gene families, dlx2a, dlx5a, hand2, col2a1a, smo, fgfr1a and fgfr1bl, are significantly downregulated in prdm1a−/− embryos (Fig. 5C). These genes are required for pectoral fin/limb induction, patterning, outgrowth, and collagen production (Ahn and Ho, 2008; Akimenko et al., 1994; Chen et al., 2001; Dale and Topczewski, 2011; Heude et al., 2014; Leerberg et al., 2019; Yan et al., 1995; Yelon et al., 2000). The most significant upregulated gene was complement factor 4b (c4b), which is part of the classical activation pathway in the immune system (Janeway et al., 2001). The paralog prdm1b was also upregulated in prdm1a−/− embryos, suggesting an attempt at genetic compensation. Gene ontology (GO) pathway enrichment analysis on genes downregulated in prdm1a−/− embryos revealed ‘anatomical structure morphogenesis’, ‘chromatin remodeling’, ‘skeletal system development’, ‘cell differentiation’ and ‘transcriptional regulation’ as the pathways most enriched in the downregulated genes (Fig. 5D). These pathways were expected given what we already know about PRDM1 as a transcription factor and master regulator of differentiation, and given the results of the cell proliferation and cell death assays presented here (Fig. S2) (reviewed in Bikoff et al., 2009).

To validate the RNA-seq data and determine the effect of prdm1a loss on gene expression, we performed real-time quantitative PCR (RT-qPCR) on the anterior half of embryos at 24 and 48 hpf and hybridization chain reaction (HCR) at 48 hpf for select genes in WT and prdm1a−/− whole embryos. (Fig. 6; Fig. S5). We first probed for prdm1a and tbx5a, an early marker of fin initiation. At 48 hpf, prdm1a is highly expressed in the fin mesenchyme and AF of WT embryos. In null mutants, prdm1a transcripts were detected throughout the fin bud and at even higher levels than those in WT, suggesting that the mRNA is not susceptible to nonsense-mediated decay at this stage (Fig. 6). The cells may be overproducing prdm1a transcripts to compensate for the loss of functional protein. tbx5a was highly expressed in the fin mesenchyme of both WT and prdm1a−/− embryos at 48 hpf with no significant difference (Fig. 6A,B). The intensity of signal in the fin mesenchyme was quantified by measuring the total cell fluorescence and correcting for area and background (corrected total cell fluorescence or CTCF). Next, as the Fgf receptor genes fgfr1a and fgfr1bl were significantly downregulated in the RNA-seq dataset (Fig. 5C), we looked at the expression of the gene encoding their ligand, fgf10a, which has been shown to be decreased in hypomorphs and morphants (Lee and Roy, 2006; Mercader et al., 2006). In WT embryos, fgf10a was highly expressed in the fin mesenchyme but significantly reduced in prdm1a−/− embryos (Fig. 6C-G; Fig. S5A,G). This was quantified by measuring the signal intensity along a line drawn from the most proximal to most distal point of the fin bud and normalizing the intensity and distance between 0 and 1 (Fulton et al., 2020). fgf10a is a marker of fin induction and is known for signaling downstream to fgf8a in the AF to regulate differentiation and outgrowth along the proximal-distal axis (Kawakami et al., 2004b). In prdm1a−/− embryos, fgf8a expression is significantly decreased, which is consistent with the observed truncated-fin phenotype (Fig. 6H-L; Fig. S5L). RNA-seq also showed that dlx5a, another marker of outgrowth, was downregulated (Fig. 5C). By HCR analysis, dlx5a was highly expressed in the fin mesenchyme and cleithrum and was co-expressed with prdm1a in AF cells (Fig. 6M). Intriguingly, dlx5a expression was decreased in the mesenchyme and cleithrum but increased in the AF of prdm1a−/− embryos (Fig. 6N-P; Fig. S5J). Taken together, these data suggest disruptions in the AF, which is analogous to the tetrapod AER (reviewed in Yano and Tamura, 2013). Misregulation in the AER is a common pathogenic feature in SHFM (reviewed in Duijf et al., 2003). Finally, we probed for sonic hedgehog (shha), the morphogen required for anterior/posterior digit patterning. In the most posterior part of the fin bud, the ZPA, there was a significant decrease of shha in prdm1a−/− embryos (Fig. 6Q,R). This was expected in that we also saw a downregulation of the receptor and Shh target smo in our RNA-seq dataset (Fig. 5C). Our gene expression results in null prdm1a−/− mutants are consistent with published studies in mice as well as morphant and hypomorph prdm1atp39/tp39 zebrafish studies (Lee and Roy, 2006; Mercader et al., 2006; Robertson et al., 2007). Taken together, these data suggest that Prdm1a is required for regulating Fgf signaling in the fin mesenchyme as well as outgrowth and anterior/posterior patterning in the AF and ZPA.

Fig. 6.

Prdm1a acts downstream of fin initiation and regulates Fgf signaling in the fin mesenchyme required for outgrowth and anterior/posterior patterning. (A,C,H,M,Q) Lateral views of pectoral fins from whole-mount WT and prdm1a−/− mutant embryos after hybridization chain reaction (HCR) was performed at 48 hpf. Scale bars: 50 µm. (A) tbx5a (pectoral fin initiation) and prdm1a expression (n=8 WT and 6 prdm1a−/− embryos). (B) Quantification of tbx5a expression using corrected total cell fluorescence (CTCF) showed no significant difference. (C) fgf10a (pectoral fin induction) and prdm1a expression (n=6 for each genotype). (D,E) Quantification of (D) fgf10a and (E) prdm1a expression along a line drawn from the most proximal to the most distal point of the fin bud using the line scan tool on ImageJ. Intensity and distance were normalized between 0 and 1. (F) Maximum normalized intensity of fgf10a shows a decrease in prdm1a−/− compared to WT embryos. (G) Length of the fin as measured by fgf10a gene expression. (H) fgf8a (AF outgrowth marker) and prdm1a expression (n=6 for each genotype). (I,J) Quantification of (I) fgf8a and (J) prdm1a expression along a line drawn from the most proximal to the most distal point of the fin bud. (K) Maximum normalized intensity of fgf8a shows a decrease in prdm1a−/− compared to WT embryos. (L) Length of the fin as measured by fgf8a gene expression. (M) dlx5a (outgrowth marker) and prdm1a expression (n=3 WT and 4 prdm1a−/− embryos). (N,O) Quantification of (N) dlx5a and (O) prdm1a expression along a line drawn from the most proximal to the most distal point of the fin bud. (P) Length of the cleithrum is decreased in prdm1a−/− compared to WT embryos (P=0.0143). (Q) shha (anterior/posterior patterning) and prdm1a expression (n=8 for each genotype). (R) Expression of shha was quantified using CTCF shows a decrease in prdm1a−/− compared to WT embryos (P=0.0294). Solid lines in line intensity graphs represent the mean±s.d. Statistical comparisons were made using unpaired, two-tailed, independent Student's t-test. All images are maximum projections of lateral views of the pectoral fin. The background was subtracted using the rolling ball feature in ImageJ (50 pixels). AF, apical fold; cl, cleithrum; CTCF, corrected total cell fluorescence; D, distal; hpf, hours post fertilization; P, proximal; ZPA, zone of polarizing activity.

Fig. 6.

Prdm1a acts downstream of fin initiation and regulates Fgf signaling in the fin mesenchyme required for outgrowth and anterior/posterior patterning. (A,C,H,M,Q) Lateral views of pectoral fins from whole-mount WT and prdm1a−/− mutant embryos after hybridization chain reaction (HCR) was performed at 48 hpf. Scale bars: 50 µm. (A) tbx5a (pectoral fin initiation) and prdm1a expression (n=8 WT and 6 prdm1a−/− embryos). (B) Quantification of tbx5a expression using corrected total cell fluorescence (CTCF) showed no significant difference. (C) fgf10a (pectoral fin induction) and prdm1a expression (n=6 for each genotype). (D,E) Quantification of (D) fgf10a and (E) prdm1a expression along a line drawn from the most proximal to the most distal point of the fin bud using the line scan tool on ImageJ. Intensity and distance were normalized between 0 and 1. (F) Maximum normalized intensity of fgf10a shows a decrease in prdm1a−/− compared to WT embryos. (G) Length of the fin as measured by fgf10a gene expression. (H) fgf8a (AF outgrowth marker) and prdm1a expression (n=6 for each genotype). (I,J) Quantification of (I) fgf8a and (J) prdm1a expression along a line drawn from the most proximal to the most distal point of the fin bud. (K) Maximum normalized intensity of fgf8a shows a decrease in prdm1a−/− compared to WT embryos. (L) Length of the fin as measured by fgf8a gene expression. (M) dlx5a (outgrowth marker) and prdm1a expression (n=3 WT and 4 prdm1a−/− embryos). (N,O) Quantification of (N) dlx5a and (O) prdm1a expression along a line drawn from the most proximal to the most distal point of the fin bud. (P) Length of the cleithrum is decreased in prdm1a−/− compared to WT embryos (P=0.0143). (Q) shha (anterior/posterior patterning) and prdm1a expression (n=8 for each genotype). (R) Expression of shha was quantified using CTCF shows a decrease in prdm1a−/− compared to WT embryos (P=0.0294). Solid lines in line intensity graphs represent the mean±s.d. Statistical comparisons were made using unpaired, two-tailed, independent Student's t-test. All images are maximum projections of lateral views of the pectoral fin. The background was subtracted using the rolling ball feature in ImageJ (50 pixels). AF, apical fold; cl, cleithrum; CTCF, corrected total cell fluorescence; D, distal; hpf, hours post fertilization; P, proximal; ZPA, zone of polarizing activity.

Prdm1a directly binds to and regulates outgrowth genes in the pectoral fin

Given that Prdm1a requires its zinc finger domain during pectoral fin development, we next asked whether Prdm1a directly binds to genes that were identified in the RNA-seq to regulate their expression. We isolated EGFP-positive pectoral fin cells from Tg(Mmu:Prx1-EGFP) WT fish at 24 hpf (Fig. 7A) and performed CUT&RUN (Meers et al., 2019; Skene and Henikoff, 2017; Ye et al., 2021). We used an IgG antibody and antibodies against histone H3 acetylated at lysine 27 (H3K27Ac) and Prdm1a (von Hofsten et al., 2008), and sequenced the samples on the Illumina NovaSEQ 6000 system. We observed 15,361 Prdm1a-occupied peaks (Fig. 7B,C; Fig. S6A,B). Of these, 29.81% were associated with promoter regions, 10.96% with introns and 58.81% with distal intergenic regions (Fig. 7D; Fig. S6C). We then subjected the Prdm1a peaks to GO pathway analysis and found enrichment for pathways involved in transcriptional regulation, such as ‘protein dimerization activity’, ‘transcription coregulator activity’ and ‘histone binding’ (Fig. 7E). We also performed motif enrichment analysis and identified a significant enrichment of Hox transcription factor-binding sites. Several Hox transcription factors are known to be required for pectoral fin development, such as Hoxd11a, Hoxa13a, Hoxa13b and Hoxa11b (Fig. 7F; Fig. S6E; Table S3) (Nakamura et al., 2016; Sordino et al., 1996, 1995). To our knowledge, Prdm1a has not yet been shown to interact with Hox transcription factors. We also mapped Prdm1a binding sites to the nearest genes and found that Prdm1a directly binds to putative enhancer and promoter regions of critical pectoral fin and limb development genes, including fgfr1a, dlx5a, dlx6a and smo (Fig. 7G; Fig. S6F). This is consistent with what was observed in our RNA-seq dataset in that the expression of these genes is significantly downregulated in prdm1a−/− compared to that in WT embryos. The data suggest that Prdm1a directly binds to the regulatory sequences of these genes and functions as an activator to regulate fin induction, outgrowth and anterior/posterior patterning.

Fig. 7.

Prdm1a directly binds to and regulates limb genes. (A) Lateral view of EGFP-positive pectoral fins from the Tg(Mmu:Prx1-EGFP) zebrafish line at 24 hpf before CUT&RUN was performed. Scale bar: 200 µm. (B,C) Coverage heatmaps of (B) H3K27Ac and (C) Prdm1a binding across the genome 1.5 kb upstream and downstream of the peak center. (D) Annotation of enriched binding sites by Prdm1a. (E) Enriched Prdm1a peaks were subjected to Gene Ontology (GO) terms analysis using ChIPseeker's enrichGO function. (F) Prdm1a peaks were subjected to motif analysis using HOMER. The top ten motifs as well as known limb-related motifs are shown along with q-values. (G) Tracks showing H3K27Ac enrichment (open chromatin) and Prdm1a-binding sites for fgfr1a, dlx5a, dlx6a and smo. There is variability between replicates, but the overall trends are comparable (see Fig. S5). e, eye; pf, pectoral fin; TF, transcription factor; y, yolk.

Fig. 7.

Prdm1a directly binds to and regulates limb genes. (A) Lateral view of EGFP-positive pectoral fins from the Tg(Mmu:Prx1-EGFP) zebrafish line at 24 hpf before CUT&RUN was performed. Scale bar: 200 µm. (B,C) Coverage heatmaps of (B) H3K27Ac and (C) Prdm1a binding across the genome 1.5 kb upstream and downstream of the peak center. (D) Annotation of enriched binding sites by Prdm1a. (E) Enriched Prdm1a peaks were subjected to Gene Ontology (GO) terms analysis using ChIPseeker's enrichGO function. (F) Prdm1a peaks were subjected to motif analysis using HOMER. The top ten motifs as well as known limb-related motifs are shown along with q-values. (G) Tracks showing H3K27Ac enrichment (open chromatin) and Prdm1a-binding sites for fgfr1a, dlx5a, dlx6a and smo. There is variability between replicates, but the overall trends are comparable (see Fig. S5). e, eye; pf, pectoral fin; TF, transcription factor; y, yolk.

Approximately 50% of SHFM cases have an unknown genetic cause. Chromosomal deletions and translocations at 6q21 have long been associated with SHFM, although a candidate gene has not yet been isolated (Braverman et al., 1993; Correa-Cerro et al., 1996; Duran-Gonzalez et al., 2007; Gurrieri et al., 1995; Hopkin et al., 1997; Pandya et al., 1995; Tsukahara et al., 1997; Viljoen and Smart, 1993). We identified three novel variants in PRDM1 in families with SHFM, which segregated with the phenotype or arose de novo. PRDM1 has been previously implicated in vertebrate limb development (Ha and Riddle, 2003; Lee and Roy, 2006; Mercader et al., 2006; Robertson et al., 2007; Vincent et al., 2005; Wilm and Solnica-Krezel, 2005) and, here, we show that PRDM1 variants likely result in SHFM and limb defects in humans. Each of the three variants is rare, negatively affects the DNA-binding zinc finger domain of the protein and is pathogenic as a heterozygous allele. We have shown through transient overexpression assays in zebrafish that the variants are pathogenic and fail to rescue cartilage elements of the pectoral fins of prdm1a−/− zebrafish embryos compared to the WT allele. Moreover, the variants likely have a dominant-negative effect in that they produce pectoral fin defects in WT embryos upon overexpression.

Using stable, conditional overexpression experiments, we define the functional domain of Prdm1a specifically during pectoral fin development. PRDM1 consists of a SET domain at its N-terminus, followed by a proline/serine-rich domain, and five zinc fingers. PRDM1 can recruit epigenetic modifiers to its domains as well as bind DNA directly to regulate transcription. We show that Prdm1a requires both its proline/serine-rich and zinc finger domains for pectoral fin morphogenesis. Deleting either domain fails to rescue the pectoral fin, whereas deletion of the SET domain rescues the cleithrum, endoskeletal disk, scapulocoracoid and postcoracoid. The SET domain of PRDM1 does not have intrinsic methyltransferase activity in vivo and has not been shown to bind with cofactors (Cheng et al., 2005; Hohenauer and Moore, 2012; Martin and Zhang, 2005). Removing the proline/serine and zinc finger domains together also fails to rescue, but it does not produce a more severe phenotype. This implies that both domains are required during pectoral fin development. Prdm1a likely directly binds to DNA with its zinc finger domain and then recruits cofactors to its proline/serine domain. If Prdm1a cannot bind, then neither can its cofactors. Likewise, binding to the DNA alone cannot repress or induce expression of that gene. Our zebrafish data highlight the importance of the proline/serine and zinc finger domains in PRDM1, which were disrupted in the SHFM families in this study. In addition, the data will be useful for predicting the pathogenicity of PRDM1 variants that may later arise.

Within the limb GRN, PRDM1 has been proposed to act downstream of retinoic acid signaling and limb initiation, and upstream of FGF signaling to induce limb formation (Lee and Roy, 2006; Mercader et al., 2006; Robertson et al., 2007). Its conserved expression in the AER of mice and chick and in the AF of zebrafish suggests an important role for PRDM1 in outgrowth (Ha and Riddle, 2003; Lee and Roy, 2006; Mercader et al., 2006; Robertson et al., 2007). However, there is no significant difference in cell proliferation or cell death in the pectoral fin with a loss of Prdm1a, suggesting that it may be involved in the differentiation of cells in the AER/AF. We performed RNA-seq on isolated pectoral fin cells and found that there was an almost equal distribution of upregulated and downregulated genes in prdm1a−/− compared to WT embryos, although all key limb genes were downregulated (Fig. 6B,C). Given that Prdm1a is traditionally considered a gene repressor, it is surprising that the genes known to be involved in pectoral fin and limb development were all downregulated in prdm1a−/− embryos, including Fgf receptors, col2a1a, dlx2a, dlx5a, smo and hoxa/hoxd genes (Ahn and Ho, 2008; Akimenko et al., 1994; Chen et al., 2001; Dale and Topczewski, 2011; Heude et al., 2014; Leerberg et al., 2019; Yan et al., 1995; Yelon et al., 2000). Using HCR, we demonstrate that prdm1a−/− pectoral fins exhibit a significant decrease in the expression of important genes, namely fgf10a, fgf8a, dlx5a and shha. We propose that during pectoral fin induction, Prdm1a promotes mesenchymal cell outgrowth, differentiation and patterning by activating the Fgf receptor fgfr1a. Binding of Fgf10a to this receptor then leads to downstream activation of fgf8a in the AF (Kawakami et al., 2004b). In chick, FGF8 initiates a positive feedback loop and maintains expression of FGF10 for sustained limb growth (Ng et al., 2002; Ohuchi et al., 1997), but this has not yet been tested in zebrafish. Shh expression in the ZPA is required for anterior/posterior patterning, and, in zebrafish, Shha has also been shown to be required for fgf8a expression in the AF (Neumann et al., 1999). In chick and mouse, FGF8 is involved in the initiation of SHH expression in the mesoderm and ZPA (Crossley et al., 1996; Lewandoski et al., 2000). It is possible that these feedback loops also occur in the zebrafish pectoral fin. Taken together, our data suggest that a loss of Prdm1a leads to disruptions in the AF owing to misregulation of Fgf and Shh signaling (Fig. 8). These changes in expression are consistent with what has been observed in both zebrafish and mice studies, although it has not yet been shown in other tetrapods (Lee and Roy, 2006; Mercader et al., 2006; Robertson et al., 2007). Our data suggest that Prdm1a is necessary for initiating induction in the fin mesenchyme and maintenance of outgrowth, differentiation and patterning genes in the AF and ZPA.

Fig. 8.

Working model of the Prdm1a gene regulatory network during pectoral fin development. Prdm1a acts downstream of pectoral fin initiation and tbx5a (Fig. 6A,B), but upstream of induction and Fgf signaling. Prdm1a directly binds to regulatory sequences of and activates fgfr1a (Fig. 7G), allowing Fgf10a to bind (Fig. 6C-G). Fgf10a then activates Fgf8a in the apical fold (Fig. 6H-L), signaling pectoral fin outgrowth and differentiation. Prdm1a also directly binds to putative enhancers and promoter regions of smo (Fig. 7G), a receptor in Shh signaling. shha is expressed in the zone of polarizing activity (Fig. 6Q,R). It is required for anterior/posterior patterning as well as regulating fgf8a expression. Dashed arrows illustrate additional feedback loops that have been demonstrated in mice and/or chick but have not yet been shown in zebrafish. Finally, Prdm1a directly binds to putative enhancers of dlx5a and dlx6a, additional fin outgrowth markers (Fig. 6M-P).

Fig. 8.

Working model of the Prdm1a gene regulatory network during pectoral fin development. Prdm1a acts downstream of pectoral fin initiation and tbx5a (Fig. 6A,B), but upstream of induction and Fgf signaling. Prdm1a directly binds to regulatory sequences of and activates fgfr1a (Fig. 7G), allowing Fgf10a to bind (Fig. 6C-G). Fgf10a then activates Fgf8a in the apical fold (Fig. 6H-L), signaling pectoral fin outgrowth and differentiation. Prdm1a also directly binds to putative enhancers and promoter regions of smo (Fig. 7G), a receptor in Shh signaling. shha is expressed in the zone of polarizing activity (Fig. 6Q,R). It is required for anterior/posterior patterning as well as regulating fgf8a expression. Dashed arrows illustrate additional feedback loops that have been demonstrated in mice and/or chick but have not yet been shown in zebrafish. Finally, Prdm1a directly binds to putative enhancers of dlx5a and dlx6a, additional fin outgrowth markers (Fig. 6M-P).

Prdm1a has traditionally been known as a repressor, and it was initially thought that Prdm1a regulates Fgf signaling and the ensuing cascade by blocking the expression of an inhibitor of fgf10a transcription (Mercader et al., 2006). However, in this study, we performed CUT&RUN on isolated pectoral fin cells and showed that Prdm1a directly binds to regulatory sequences of its receptor, fgfr1a, suggesting that it is activating Fgf signaling. We also show that Prdm1a directly binds to putative enhancers or promoter regions of smo and activates anterior/posterior patterning. Its role as an activator is uncommon, although not novel. We have previously shown that Prdm1a directly activates genes, such as tfap2a and foxd3, during zebrafish neural specification (Powell et al., 2013). More recently, it has been shown to interact with Kdm4a, a histone demethylase, to activate chick neural, neural crest (NC) and sensory specification genes (Prajapati et al., 2019). Given that its traditional cofactors are repressors, i.e. HDAC1/2, Groucho proteins, LSD1 and Prmt5 (Ancelin et al., 2006; Ren et al., 1999; Su et al., 2009; Yu et al., 2000), it is unlikely that Prdm1a binds with these factors during the activation of these particular limb genes. However, ‘chromatin modification’ was one of the most downregulated pathways in prdm1a−/− embryos from our GO pathway enrichment analysis of the RNA-seq dataset (Fig. 5D), and ‘transcription coregulator activity’ was enriched in the CUT&RUN dataset (Fig. 7E). We hypothesize that Prdm1a may act with Kdm4a to directly activate fgfr1a and smo, although additional experiments are needed to determine this or uncover other cofactors. Motif analysis of Prdm1a-bound peaks predicts enrichment of Hox transcription factor motifs, including Hoxd11a, Hoxa13a, Hoxa13b and Hoxa11b, that are known to be required for pectoral fin development (Fig. 7F) (Nakamura et al., 2016; Sordino et al., 1996, 1995).

We have shown in zebrafish that Prdm1a is required for the maintenance of outgrowth, differentiation and patterning genes in the AF, an analogous structure to the tetrapod AER. Importantly, disruptions to the AER are considered the primary disease mechanisms of SHFM (reviewed in Duijf et al., 2003). For example, Dlx5 and Dlx6, homeodomain transcription factors causing SHFM type I (Crackower et al., 1996; Scherer et al., 1994; Ullah et al., 2017; Wang et al., 2014), are required for maintaining proliferation in medial cells of the AER (Robledo et al., 2002). TP63 (SHFM type IV) regulates formation and differentiation of the AER, and regulates ectodermal development (Mills et al., 1999; Yang et al., 1999). Deletions and chromosomal rearrangements at the DLX5/DLX6 locus on 7q21-q22 and variants in TP63 are the two most common genetic causes for SHFM (Crackower et al., 1996) (reviewed in Sowinska-Seidler et al. (2014)). Interestingly, TP63 has also been shown to act as an upstream regulator of Dlx5/Dlx6 by binding to cis-regulatory elements near the promoter region (Kouwenhoven et al., 2010; Lo Iacono et al., 2008). Although the SHFM families in this study tested normally for TP63, it is possible that the PRDM1 variants disrupted part of this GRN as well as the AER of the developing limb bud.

Indeed, one of the more interesting genes identified by our transcriptomic and CUT&RUN analyses was dlx5a. In our HCR assays, we showed that at 48 hpf, prdm1a−/− embryos exhibited decreased expression of dlx5a in the mesenchyme, but increased expression in the AF where prdm1a is also co-expressed (Fig. 6M,N). Given that DLX5 mutations lead to SHFM type I (MIM #183600) (Crackower et al., 1996; Scherer et al., 1994; Ullah et al., 2017; Wang et al., 2014) and that dlx5a is directly regulated by Prdm1a, we asked whether there is a genetic interaction between the two genes. We crossed a hypomorphic allele dlx5aj1073Et [referred to as Tg(dlx5a:EGFP)] (Talbot et al., 2010) to prdm1a+/− fish, intercrossed the double-heterozygous animals and performed cartilage staining on the resulting larvae at 4 dpf. dlx5a morphants have pectoral fin defects that vary in severity (Heude et al., 2014). Interestingly, Tg(dlx5a:EGFP) heterozygotes and homozygotes did not have overt pectoral fin defects at this stage (Fig. S7C,E). When combined with prdm1a−/−, we found that the length of the endoskeletal disk and the area of the scapulocoracoid and postcoracoid were slightly increased and trending towards a rescue; however, the results were not significant (Fig. S7D,F,H,J). The partial loss of Dlx5a could have helped balance the high expression in the AF of prdm1a−/− embryos and rescued pectoral fin outgrowth, but, because it is a hypomorph, it may have been too weak to produce a more drastic rescue.

In addition, expression of dlx5a in the cleithrum was significantly decreased in prdm1a−/− embryos (Fig. 6M,O). The cleithrum is part of the shoulder girdle in bony fishes. It is located at the border between the NC-derived pharyngeal arches, which give rise to the craniofacial skeleton and the mesodermal pectoral fin. Because of its position, some have hypothesized that, like the clavicle in mammals, the cleithrum may be composed of both NC and mesodermal cells (Matsuoka et al., 2005). This could have important implications in human disease in that craniofacial and limb defects often co-occur, including in SHFM (reviewed in Gurrieri and Everman, 2013; Truong and Artinger, 2021). Although there is currently no evidence that NC cells contribute to the cleithrum (Kague et al., 2012), NC cells have been labeled in gill pillar cells of zebrafish (Mongera et al., 2013) as well as the posterior gill arches of the little skate (Leucoraja erinacea) (Sleight and Gillis, 2020). The gill arches are hypothesized to give rise to paired fins in jawed vertebrates, implying a serial homology between the two structures (Gegenbaur, 1878; Sleight and Gillis, 2020). Given the importance of Prdm1a in the pharyngeal arches and now the pectoral fin, it is interesting to speculate whether the two structures are connected (Artinger et al., 1999; Birkholz et al., 2009; Ha and Riddle, 2003; Lee and Roy, 2006; Mercader et al., 2006; Robertson et al., 2007; Roy and Ng, 2004; Vincent et al., 2005; Wilm and Solnica-Krezel, 2005).

In summary, we have identified novel variants in PRDM1 that result in SHFM phenotypes and limb defects with incomplete penetrance and variable expressivity in humans. Variants affecting the ability of the protein to recruit cofactors and bind to DNA are detrimental for proper limb formation. Moreover, we demonstrate that a loss of Prdm1a leads to disruptions in the AF of zebrafish pectoral fins. We show that Prdm1a directly binds to putative enhancer and promoter regions of fgfr1a, dlx5a, dlx6a and smo during fin development, and its ability to do so is critical for proper outgrowth, differentiation and patterning (Fig. 8). Although zebrafish fin and tetrapod limb development are distinct from one another, our study provides important clues into the potentially pathogenic role of PRDM1 in human limb development and improves our understanding of the limb GRN.

Zebrafish husbandry

Zebrafish were maintained as previously described (Westerfield, 2000). The WT strain used was AB (Zebrafish International Resource Center) and the mutant lines used were prdm1am805 (nrd; referred to as prdm1a−/−) (Artinger et al., 1999; Hernandez-Lagunas et al., 2005) and dlx5aj1073Et [referred to as Tg(dlx5a:EGFP)] (Talbot et al., 2010). Embryos were staged following previously published standards (Kimmel et al., 1995). All experiments were reviewed and approved by the Institutional Animal Care and Use Committee (IACUC) at the University of Colorado Denver Anschutz Medical Campus (IACUC protocol #147) and conform to the National Institutes of Health regulatory standards of care and treatment.

Participants in SHFM study

SHFM individuals were seen in the clinic due to a history of non-syndromic limb and digit malformations. X-ray and pedigree analyses indicated a diagnosis of non-syndromic SHFM with dominant inheritance but variable penetrance for each. In addition to testing for common variants associated with SHFM, standard chromosomal karyotyping and microarray were performed but did not reveal any abnormalities. In SHFM family #1 (PRDM1c.712_713insT), DNA derived from whole blood was used to perform WES, which identified variants in the PRDM1 gene. Targeted sequencing for PRDM1 was then performed on an additional 75 unrelated SHFM individuals seen at the clinic, which identified two additional variants. Ascertainment of human subjects, samples and data was reviewed and approved by the Greenwood Genetics Center Self Regional Healthcare Institutional Review Board (IRB) (approval #33). Informed consent to be included in the study and be published was obtained from all subjects.

DNA extraction, exome sequencing and analysis

Exome sequencing was performed at the University of Washington Center for Mendelian Genomics (UW-CMG). Briefly, library construction and exome capture were done using an automated 96-well plate format (Perkin-Elmer Janus II). Approximately 500 ng of genomic DNA was subjected to a series of shotgun library construction steps, including fragmentation through acoustic sonication (Covaris), end polishing and A-tailing, ligation of sequencing adaptors, and PCR amplification with dual 8 bp barcodes for multiplexing. Libraries underwent exome capture using the Roche/Nimblegen SeqCap EZ v2.0 (∼36.5 Mb target). Prior to sequencing, the library concentration was determined by fluorometric assay and molecular mass distributions verified on the Agilent Bioanalyzer (consistently 150±15 bp). Barcoded exome libraries were pooled using liquid handling robotics prior to clustering (Illumina cBot) and loading. Massively parallel sequencing-by-synthesis with fluorescently labeled, reversibly terminating nucleotides was carried out on the HiSeq sequencer (Illumina). Variant detection and genotyping were performed using the HaplotypeCaller tool (https://gatk.broadinstitute.org/hc/en-us/articles/360037225632-HaplotypeCaller) from the Genome analysis Toolkit (GATK) (v3.7). Variant data for each sample were formatted [variant call format (VCF)] as ‘raw’ calls that contain individual genotype data for one or multiple samples and flagged using the filtration walker (GATK) to mark sites that were of lower quality or were false positives [e.g. low-quality scores (Q50), allelic imbalance (ABHet 0.75), long homopolymer runs (HRun>3) and/or low quality by depth (QD<5)].

Sample identity and relationships were confirmed by sex and pedigree checks implemented in PLINK v1.90 (Chang et al., 2015) and KING v1.4.0 (Manichaikul et al., 2010). We extracted single-nucleotide variants (SNVs) and short insertions/deletions (indels) with per-sample read depth between 2 and 500, per-sample minimum genotype quality of 20, and minimum alternate allele count of 2 using bcftools v1.2 (Danecek et al., 2021), yielding 28,946 variants. These variants were annotated using the Variant Effect Predictor (VEP v75) (McLaren et al., 2016) and loaded into a database using GEMINI (v0.14.1) (Paila et al., 2013). We extracted 57 variants segregating in an autosomal-dominant pattern that had reference alternate allele frequencies (AAF) <0.01 [1000 genomes, National Heart, Lung, and Blood Institute (NHLBI) GO Exome Sequencing Project (ESP), Exome Aggregation Consortium (EXaC); The 1000 Genomes Project Consortium, 2015]. Ten of these variants had AAF<0.001 and were predicted to be functional (e.g. missense); each of these variants were confirmed in Integrative Genomics Viewer (Robinson et al., 2011). We performed a literature review of the genes implicated by these ten variants, and only PRDM1 had evidence for a role in limb development. Further selection was performed by selecting rare SNVs that were considered damaging by at least two bioinformatics tools [including Polymorphism Phenotyping v2 HumDiv (PP2_HD), MutationTaster, Sorting Intolerant From Tolerant (SIFT) and Combined Annotation-Dependent Depletion (CADD)] in dbNSFP (Liu et al., 2016). For indels, bioinformatics analysis was performed using MutationTaster (Schwarz et al., 2010).

SHFM Sanger sequencing

PRDM1 variants were confirmed by Sanger sequencing using BigDye Terminator v3.1 Cycle Sequencing kit (Thermo Fisher Scientific) on an ABI3100 automatic DNA analyzer (Applied Biosystems) following the manufacturer's instructions. The alignment and analysis of the sequences were done using the DNASTAR program (Lasergene).

Alcian Blue cartilage staining

Zebrafish were stained for cartilage as previously described (Walker and Kimmel, 2007). In short, 4 dpf larvae were fixed in 2% paraformaldehyde (PFA) at room temperature for 1 h. Larvae were then washed in 100 mM Tris (pH 7.5)/10 mM MgCl2 before rocking overnight at room temperature in Alcian Blue stain (pH 7.5) [0.04% Alcian Blue, 80% ethanol, 100 mM Tris (pH 7.5) and 10 mM MgCl2]. Larvae were destained and rehydrated in a series of ethanol washes (80%, 50% and 25%) containing 100 mM Tris (pH 7.5) and 10 mM MgCl2, and then bleached for 10 min in 3% H2O2/0.5% KOH. Finally, larvae were rinsed twice in 25% glycerol/0.1% KOH to remove the bleach and stored at 4°C in 50% glycerol/0.1% KOH. The pectoral fins of stained larvae were dissected, flat mounted in 50% glycerol/0.1% KOH, and imaged on an Olympus BX51 WI microscope. Measurements of the pectoral fin were performed in an anonymized manner in ImageJ, averaged for each individual, and then compared using a one-way ANOVA followed by a Tukey's post hoc test relative to uninjected prdm1a−/− mutants. Sample size refers to the number of individuals and is included in the figure legends.

Immunofluorescence

Zebrafish embryos were collected at the indicated time points and fixed in 4% PFA at room temperature for 1 h. Embryos were washed twice in 1× PBS (pH 7.3), dehydrated and permeabilized in two 10-min washes in 100% methanol at room temperature. Embryos were stored for at least 24 h at −20°C in fresh methanol. A graded series of methanol in PBS containing 0.01% Tween-20 (PBST) solutions was used to rehydrate the embryos (75%, 50%, 25% and 0%). The embryos were then equilibrated in 150 mM Tris (pH 9.5) at room temperature for 5 min and then 70°C for 20 min before being washed several times in PBST and then distilled water for 5 min. Embryos were then incubated at room temperature in blocking solution [2% goat serum, 1% bovine serum albumin (BSA), 1% dimethyl sulfoxide (DMSO) and 0.1% Triton X-100 in 1× PBS] before adding the primary antibody diluted in blocking solution. The antibodies used were anti-phosphoH3 (1:500, Sigma-Aldrich, H0412, lot #088M4842V) and anti-cleaved caspase 3 (1:500, Cell Signaling Technology, 9661, lot #47). The embryos were incubated overnight at 4°C. Following primary antibody incubation, samples were washed thoroughly in PBS with 0.1% Triton X-100, then incubated overnight in fluorescently tagged goat anti-rabbit Alexa Fluor 594 secondary antibody (1:500, Invitrogen, A-11012) at 4°C. Embryos were rinsed in PBS with 0.1% Triton X-100 before adding DAPI diluted in PBS for 1 h. Embryos were rinsed and stored at 4°C. Whole embryos were mounted in 0.2% low-melt agarose and imaged on a Leica SP8 confocal microscope at 10× and 40× magnification. Sample size refers to the number of individuals and is included in the figure legends. Quantification of cell numbers were completed on maximum projections of z-stack images using ImageJ. Counts were compared using an unpaired, two-tailed independent t-test.

mRNA overexpression in zebrafish

hPRDM1 variant cDNA was synthesized into a pCS2+ backbone (Rupp et al., 1994; Turner and Weintraub, 1994) using Gateway cloning. cDNA was linearized and transcribed using the mMessage mMachine T7 Transcription Kit (Thermo Fisher Scientific). prdm1a+/− fish were intercrossed and the different hPRDM1 mRNA variants (diluted 1:10 in water and Phenol Red) were injected into resulting embryos at the single-cell stage. Embryos were staged throughout the first four days of development to ensure that there was no developmental delay or unassociated pathologies due to the mRNA overexpression. At 4 dpf, larvae were collected for Alcian Blue staining. Sample size refers to the number of individuals and is included in the figure legends.

HCR v3.0 and quantification

Probes for prdm1a, fgf10a, fgf8a, dlx5a, tbx5a and shh were purchased from Molecular Instruments (https://www.molecularinstruments.com/). Whole-mount HCR was performed according to the manufacturer's instructions with minor modifications (Choi et al., 2016, 2018). Embryos were fixed overnight at 4°C in 4% PFA, washed in PBS, and dehydrated and permeabilized in two 10-min washes in 100% methanol at room temperature. Embryos were stored for at least 24 h at −20°C in fresh methanol. A graded series of methanol /PBST solutions was used to rehydrate the embryos (75%, 50%, 25% and 0%). Embryos were then treated with proteinase K (10 μg/ml) for 5 min (24 hpf) or 15 min (48 hpf), washed twice in PBST, fixed for 20 min in 4% PFA, and then washed five times in PBST. Following hybridization with the probe solution, the probes were saved and stored at −20°C for future use. Likewise, following the amplification stage, hairpins were saved and stored at −20°C. Recycled hairpins were heated to 95°C for 90 s and cooled (Hybridization Chain Reaction (HCR) In Situ Protocol; https://dx.doi.org/10.17504/protocols.io.bunznvf6). Embryos were stored in PBS at 4°C protected from light. Whole embryos were mounted in 0.2% low-melt agarose and imaged on a Leica SP8 confocal microscope at 10× and 40× magnification. Embryos were then genotyped following the protocol in Rossi et al. (2009) with slight modifications. Following DNA extraction, PCR was performed in M buffer [2 mM MgCl2, 13.7 mM Tris-HCl (pH 8.4), 68.4 mM KCl, 0.001% gelatin, 1.8 mg/ml protease-free BSA and 136 μM each dATP/CTP/GTP/TTP] with GoTaq Flexi DNA polymerase (Promega) and digested overnight in Fok1 enzyme (New England Biolabs) at 37°C.

HCR images were first processed by performing the ‘rolling ball’ background subtraction (50 pixels) on the sum slice projection in ImageJ (Sternberg, 1983). The signal intensity for tbx5a and shha was quantified by calculating the CTCF. The average fluorescence intensity was calculated as:
formula

An independent, unpaired two-tailed t-test was performed to compare the CTCF and the area of expression in WT/heterozygotes compared to those in prdm1a−/− mutants. The signal intensity for fgf10a, fgf8a, dlx5a and prdm1a was quantified using the line tool in ImageJ as previously described (Fulton et al., 2020). A line was drawn from the most proximal end of the pectoral fin to the most distal tip. The signal intensity along the 30-pixel-wide line was measured. The length of the fin was normalized between 0 and 1, with 0 representing the most proximal end and 1 being distal, for each individual. The signal intensity was normalized for each gene by calculating a z-score: , where x is the raw intensity for a single sample, and ‘min’ and ‘max’ are the minimum and maximum intensity, respectively, among all samples for that gene. The average signal intensity along the line, the maximum intensity and length of expression are shown. The definition of n varies for each experiment and is specified in the figure legends. Figures are three-dimensional maximum projections of lateral views of the pectoral fin. The background was subtracted using the ‘rolling ball’ function in ImageJ (50 pixels). Sample size refers to the number of individuals and is included in the figure legends.

RNA isolation for RT-qPCR

Total RNA was isolated from pooled WT/heterozygous and prdm1a−/− mutant embryos after removal of the trunk at 24 and 48 hpf with TRIzol reagent (Invitrogen) and phenol/chloroform (5-10 embryo heads per biological replicate). RNA (500 ng) was reverse transcribed to cDNA with SuperScript III First-Strand Synthesis System (Invitrogen) for RT-qPCR. Taqman primers for prdm1a, fgf10a, fgf8a, dlx5a, dlx6a, tp63, fgf24 and b-actin (actb1) were purchased from Thermo Fisher Scientific. b-actin was used as the zebrafish internal control. Reactions were performed in at least three biological and technical replicates. Transcript abundance and relative fold change were quantified using the 2−ΔΔCt method relative to control. Relative expression was compared using an unpaired, two-tailed independent t-test. P-values less than 0.05 were considered statistically significant.

Molecular cloning

A full-length open reading frame of prdm1a was amplified from cDNA as previously described (Hernandez-Lagunas et al., 2005). Amplicons were gel extracted using the Zymoclean Gel DNA Recovery Kit (Zymo Research) and recombined into the pDONR221 plasmid (Kwan et al., 2007) using BP Clonase II following the manufacturer's instructions (Invitrogen) (Kwan et al., 2007) to make the pME-prdm1a construct. Sequences were confirmed with Sanger sequencing.

To delete the different protein domains of Prdm1a, unique restriction-digest sites were introduced into the pME-prdm1a construct using the QuikChange Lightning Multi Site-Directed Mutagenesis Kit (Agilent Technologies). Primer sequences were designed using the QuikChange Primer Design Program (https://www.agilent.com/store/primerDesignProgram.jsp) and are included in Table S2. Domain deletions were modeled after in vitro studies described previously (Ancelin et al., 2006; Gyory et al., 2004; Su et al., 2009; Yu et al., 2000). XhoI and SalI sites were introduced to flank the SET domain; AatII sites surrounded the proline/serine-rich domain; XbaI sites flanked the zinc finger domain; and XbaI sites flanked the proline/serine-rich domain and zinc finger domain. Following an overnight restriction digest at 37°C with the proper enzyme, fragments were run on a 1% agarose gel, extracted and ligated without the deleted Prdm1a domain using T4 DNA ligase (New England Biolabs) overnight at 16°C. The enzyme was inactivated at 65°C for 10 min before 2.5 μl of the reaction was transformed into Escherichia coli DH5α cells. Sequences were confirmed by Sanger sequencing. If sequences were out of frame, additional nucleotides were reinserted/deleted using the QuikChange Lightning kit and then resequenced (Table S2).

To generate Tg(hsp70l:gal4FF), p5E-hsp70l (a gift from Dr Brian Ciruna, University of Toronto, Toronto, ON, Canada), pME-gal4FF (Asakawa et al., 2008), and p3E-polyA (Kwan et al., 2007) were recombined into a pDestTol2CG2 destination vector (Kwan et al., 2007) using Gateway LR Clonase II (Invitrogen) (Kwan et al., 2007). The UAS constructs were created by recombining p5E-4XnrUAS (a gift from Dr Bruce Appel, University of Colorado Denver Anschutz Medical Campus, Aurora, CO, USA), pME-prdm1a variations and p3E-2a-EGFP (a gift from Dr Bruce Appel) into pDestTol2pA2 (Kwan et al., 2007) using LR Clonase II (Akitake et al., 2011; Kwan et al., 2007). Sequences were confirmed with Sanger sequencing.

Transgenesis

Transposase mRNA was synthesized as previously described (Kawakami et al., 2004a). To generate the Tg(hsp70l:gal4FF)co1025Tg fish, embryos were injected at the single-cell stage with 37.5 pg of the transgene, 28.7 pg of Tol2 mRNA and 150 mM KCl. Embryos were screened for EGFP expression in the heart at 24-72 hpf, grown to adulthood, and outcrossed to prdm1a+/− fish to generate stable F1 lines. Two independent Tg(hsp70l:gal4FF); prdm1a+/− lines were maintained. These lines were incrossed for microinjections to generate prdm1a−/− mutants.

The Tol2 plasmid for generation of the Tg(Mmu:Prx1-EGFP)co1026Tg fish that label the pectoral fin with EGFP was a generous gift from Dr Koji Tamura (Tohoku University, Sendai, Japan) (Hernández-Vega and Minguillón, 2011; Yano and Tamura, 2013). Embryos were injected at the single-cell stage with 60 pg of transgene, 28.7 pg of Tol2 mRNA, and 150 mM KCl. Embryos were screened for EGFP expression at 24-72 hpf, grown to adulthood, and outcrossed to WT to generate stable F1 lines. Two independent Tg(Mmu:Prx1-EGFP) lines with similar expression were maintained.

Global heat-shock experiments

Tg(hsp70l:gal4FF);prdm1a+/− fish were intercrossed and injected at the single-cell stage with 75 pg of the 4XnrUAS construct, 19.1 pg of Tol2 mRNA and 150 mM KCl. Following microinjection, embryos were heat shocked at 37°C for 60 min at 6 hpf. Embryos were returned to the incubator at 28.5°C to recover overnight. Embryos were then screened for mosaic EGFP expression at 24 hpf (Fig. 6B-E).

FACS

EGFP-positive pectoral fin cells were isolated from zebrafish embryos using FACS. Tg(Mmu:Prx1-EGFP) embryos were collected at 24 hpf (CUT&RUN) and 48 hpf (RNA-seq) and digested in Pronase (1 mg/ml) (Roche) for 5-6 min to remove the chorion. Embryos were pooled together, washed in DPBS (Gibco), and dissociated in Accumax (Innovative Cell Technologies) and DNase I (50 U/100 embryos) (Roche) at 31°C for 1.5 h. Cells were homogenized every 8 min by pipetting up and down using pipette tips decreasing in size. Cells were washed in solution (300 U DNase I in 4 ml DPBS) before filtering through a 70 μm nylon mesh cell strainer (Thermo Fisher Scientific) into 50 ml conical tubes pre-coated with 5% fetal bovine serum (FBS; Invitrogen). Cells were spun down, resuspended in basic sorting buffer [1 mM EDTA, 25 mM HEPES (pH 7.0), 1% FBS in DPBS], stained with DAPI (1:1000), and sorted using FACS at the University of Colorado Cancer Center Flow Cytometry Shared Resource (Aurora, CO, USA) on the MoFlo XDP100 sorter (Beckman Coulter) with a 100 μm nozzle tip (Beckman Coulter).

CUT&RUN

Following FACS, CUT&RUN was performed on 150,000+ sorted EGFP-positive pectoral fin cells at 24 hpf as previously described (Shull et al., 2022; Skene and Henikoff, 2017). Briefly, cells were incubated on activated concanavalin A-conjugated paramagnetic beads (EpiCypher) at room temperature for 10 min. Cells were washed in antibody buffer [20 mM HEPES (pH 7.5), 150 mM NaCl, 0.5 mM spermidine (Invitrogen), 1× Complete-Mini Protease Inhibitor tablet (Roche Diagnostics), 0.01% digitonin (Sigma-Aldrich) and 2 mM EDTA] and incubated overnight at 4°C with rotation in the respective antibody [IgG (1:100; Jackson ImmunoResearch, 111-005-003, RRID: AB_2337913), anti-H3K27ac (1:66; Cell Signaling Technology, 4353S, RRID: AB10545273) and anti-Prdm1a (1:33; rabbit polyclonal antibody from Dr Phillip Ingham, Lee Kong Chian School of Medicine, Singapore) (von Hofsten et al., 2008); validated for chromatin immunoprecipitation by Powell et al., 2013)]. Excess antibody was removed by washing in ice-cold digitonin Buffer [20 mM HEPES (pH 7.5), 150 mM NaCl, 0.5 mM spermidine and 1× Complete-Mini Protease Inhibitor tablet]. Cells were then incubated with pAG-MNase (EpiCypher) for 10 min at room temperature and washed with digitonin Buffer. Cells were rotated in 100 mM CaCl2 at 4°C for 2 h before the stop buffer (340 mM NaCl, 20 mM EDTA, 4 mM EGTA, 50 μg/ml RNaseA and 50 μg/ml glycogen) was added for 10 min at 37°C without the E. coli spike in. DNA fragments were purified with a DNA Clean & Concentrate Kit (Zymo Research). Eluted DNA fragments were amplified using the NEBNext Ultra II DNA Library Prep Kit for Illumina (New England Biolabs) following the manufacturer's instructions. Amplification of DNA was performed following guidelines outlined by EpiCypher: 98°C for 45 s; 98°C for 15 s and 60°C for 10 s, 14 cycles; and 72°C for 1 min. Samples were subjected to paired-end 150 bp sequencing on the Illumina NovaSEQ 6000 system at Novogene Corporation (Sacramento, CA, USA). CUT&RUN experiments were performed in duplicate for two biological replicates.

RNA-seq

About 250 Tg(Mmu:Prx1-EGFP);prdm1a−/− (sorted by pigment phenotype) (Artinger et al., 1999; Hernandez-Lagunas et al., 2005) and Tg(Mmu:Prx1-EGFP) WT embryos were dissected at 48 hpf to remove the brain before FACS. RNA from sorted cells was extracted using the RNAqueous-Micro Total RNA Isolation Kit (Thermo Fisher Scientific) following the manufacturer's instructions for cultured cells. DNase treatment was performed. A library was prepared using the NEBNext Ultra II Directional RNA Library Kit for Illumina following the manufacturer's instructions. Samples were subjected to sequencing on the Illumina NovaSEQ 6000 system at Novogene Corporation at a depth of over 20 million reads per sample. RNA-seq experiments were performed in duplicate for two biological replicates per genotype.

Bioinformatics analysis

CUT&RUN

Analysis was adapted from Ye et al. (2021). Following sequencing, paired reads were trimmed using Cutadapt (Martin, 2011). Trimmed reads were aligned to the zebrafish genome (danRer11) using Bowtie2 version 2.4.5 with the following options: --end-to-end --very-sensitive --no-mixed --no-discordant --no-unal (Langmead and Salzberg, 2012). Peak calling was performed using MACS2 v2.2.7.1 using the default settings (Zhang et al., 2008), and heatmaps, bigwig tracks and other statistics were generated with deepTools (Ramírez et al., 2014). Motif enrichment analysis was performed on peak files (bed files) using HOMER v4.11 (Heinz et al., 2010) and the findMotifsGenome.pl script. Called peaks were annotated and subjected to GO term analysis using the ChIPseeker R package with the enrichGO function (Wu et al., 2021; Yu et al., 2015). Replicates were analyzed separately. There was variability between the two replicates, but they were comparable and showed similar trends.

RNA-seq

Following sequencing, paired reads were trimmed and mapped to the zebrafish genome (danRer11) assembly using Spliced Transcripts alignment to a Reference (STAR) v2.7.10b (Dobin et al., 2013). Aligned counts per gene were calculated using featureCounts (Liao et al., 2014). Differential expression between WT and prdm1a−/− embryos was calculated using the DESeq2 package (Love et al., 2014). The top 250 differentially expressed genes by adjusted P-value (Padj) were plotted onto a heatmap using the pheatmaps R package (https://cran.r-project.org/web/packages/pheatmap/index.html). Gene lists were analyzed for functional annotation using GO enrichment analysis based on the PANTHER Classification System (Ashburner et al., 2000; Gene Ontology, 2021; Mi et al., 2019).

We thank members of the K.B.A laboratory for project feedback; Dr Chris Johnson for making the pME-prdm1a construct; Dr Koji Tamura for the Mmu:Prx1-EGFP plasmid; Dr Brian Ciruna for the p5E-hsp70l construct; the lab of Dr David Clouthier for cloning advice; the lab of Dr Jamie Nichols for providing the Tg(dlx5a:EGFP) line; the lab of Dr Bruce Appel for various Tol2 constructs; Dr Lee Niswander for critically reviewing the manuscript; Christine Archer and the zebrafish fish facility team for excellent animal care; Dr Dmitry Baturin with the University of Colorado Cancer Center Flow Cytometry Shared Resource for help with FACS; Dr Philip Ingham for the anti-Prdm1a polyclonal antibody; and the SHFM families for participation in the study. Exome sequencing by the University of Washington Center for Mendelian Genomics (UW-CMG) was funded by the National Human Genome Research Institute (NHGRI) and the National Heart, Lung, and Blood Institute (NHLBI) (UM1 HG006493 and U24 HG008956). This work is dedicated to the memory of Ethan Francis Schwartz, 1996-1998.

Author contributions

Conceptualization: B.T.T., C.E.S., H.F.-S., K.B.A.; Methodology: B.T.T., L.C.S.; Formal analysis: B.T.T., L.C.S., E.L., E.E.B.; Investigation: B.T.T., L.C.S.; Resources: E.G.B., M.F., C.S., D.E., C.E.S., H.F.-S., K.B.A.; Writing - original draft: B.T.T., K.B.A.; Writing - review & editing: L.C.S., E.L., E.G.B., M.F., E.E.B., M.J.B., C.E.S., H.F.-S.; Supervision: K.B.A.; Funding acquisition: B.T.T., M.J.B., C.E.S., H.F.-S., K.B.A.

Funding

This work is supported by the Eunice Kennedy Shriver National Institute of Child Health and Human Development (NICHD) (1F31HD103368 to B.T.T. and 1R03HD096320-01A1 to K.B.A.) and, in part, by the South Carolina Department of Disabilities and Special Needs. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health. Open Access funding provided by the University of Colorado. Deposited in PMC for immediate release.

Data availability

The RNA-seq and CUT&RUN data have been deposited in NCBI's Gene Expression Omnibus and are accessible through accession number GSE217486.

Agarwal
,
P.
,
Wylie
,
J. N.
,
Galceran
,
J.
,
Arkhitko
,
O.
,
Li
,
C.
,
Deng
,
C.
,
Grosschedl
,
R.
and
Bruneau
,
B. G.
(
2003
).
Tbx5 is essential for forelimb bud initiation following patterning of the limb field in the mouse embryo
.
Development
130
,
623
-
633
.
Ahn
,
D.
and
Ho
,
R. K.
(
2008
).
Tri-phasic expression of posterior Hox genes during development of pectoral fins in zebrafish: implications for the evolution of vertebrate paired appendages
.
Dev. Biol.
322
,
220
-
233
.
Akimenko
,
M. A.
,
Ekker
,
M.
,
Wegner
,
J.
,
Lin
,
W.
and
Westerfield
,
M.
(
1994
).
Combinatorial expression of three zebrafish genes related to distal-less: part of a homeobox gene code for the head
.
J. Neurosci.
14
,
3475
-
3486
.
Akitake
,
C. M.
,
Macurak
,
M.
,
Halpern
,
M. E.
and
Goll
,
M. G.
(
2011
).
Transgenerational analysis of transcriptional silencing in zebrafish
.
Dev. Biol.
352
,
191
-
201
.
Ancelin
,
K.
,
Lange
,
U. C.
,
Hajkova
,
P.
,
Schneider
,
R.
,
Bannister
,
A. J.
,
Kouzarides
,
T.
and
Surani
,
M. A.
(
2006
).
Blimp1 associates with Prmt5 and directs histone arginine methylation in mouse germ cells
.
Nat. Cell Biol.
8
,
623
-
630
.
Artinger
,
K. B.
,
Chitnis
,
A. B.
,
Mercola
,
M.
and
Driever
,
W.
(
1999
).
Zebrafish narrowminded suggests a genetic link between formation of neural crest and primary sensory neurons
.
Development
126
,
3969
-
3979
.
Asakawa
,
K.
,
Suster
,
M. L.
,
Mizusawa
,
K.
,
Nagayoshi
,
S.
,
Kotani
,
T.
,
Urasaki
,
A.
,
Kishimoto
,
Y.
,
Hibi
,
M.
and
Kawakami
,
K.
(
2008
).
Genetic dissection of neural circuits by Tol2 transposon-mediated Gal4 gene and enhancer trapping in zebrafish
.
Proc. Natl. Acad. Sci. USA
105
,
1255
-
1260
.
Ashburner
,
M.
,
Ball
,
C. A.
,
Blake
,
J. A.
,
Botstein
,
D.
,
Butler
,
H.
,
Cherry
,
J. M.
,
Davis
,
A. P.
,
Dolinski
,
K.
,
Dwight
,
S. S.
,
Eppig
,
J. T.
et al. 
(
2000
).
Gene ontology: tool for the unification of biology. The Gene Ontology Consortium
.
Nat. Genet.
25
,
25
-
29
.
Bakkers
,
J.
,
Hild
,
M.
,
Kramer
,
C.
,
Furutani-Seiki
,
M.
and
Hammerschmidt
,
M.
(
2002
).
Zebrafish ΔNp63 is a direct target of bmp signaling and encodes a transcriptional repressor blocking neural specification in the ventral ectoderm
.
Dev. Cell
2
,
617
-
627
.
Baxendale
,
S.
,
Davison
,
C.
,
Muxworthy
,
C.
,
Wolff
,
C.
,
Ingham
,
P. W.
and
Roy
,
S.
(
2004
).
The B-cell maturation factor Blimp-1 specifies vertebrate slow-twitch muscle fiber identity in response to Hedgehog signaling
.
Nat. Genet.
36
,
88
-
93
.
Bikoff
,
E. K.
,
Morgan
,
M. A.
and
Robertson
,
E. J.
(
2009
).
An expanding job description for Blimp-1/PRDM1
.
Curr. Opin. Genet. Dev.
19
,
379
-
385
.
Birkholz
,
D. A.
,
Olesnicky Killian
,
E. C.
,
George
,
K. M.
and
Artinger
,
K. B.
(
2009
).
Prdm1a is necessary for posterior pharyngeal arch development in zebrafish
.
Dev. Dyn.
238
,
2575
-
2587
.
Braverman
,
N.
,
Kline
,
A.
and
Pyeritz
,
R.
(
1993
).
Interstitial deletion of 6q associated with ectrodactyly
.
Am. J. Hum. Genet.
53
,
410
.
Castillo-Fernandez
,
J. E.
,
Spector
,
T. D.
and
Bell
,
J. T.
(
2014
).
Epigenetics of discordant monozygotic twins: implications for disease
.
Genome Med.
6
,
60
.
Chang
,
C. C.
,
Chow
,
C. C.
,
Tellier
,
L. C.
,
Vattikuti
,
S.
,
Purcell
,
S. M.
and
Lee
,
J. J.
(
2015
).
Second-generation PLINK: rising to the challenge of larger and richer datasets
.
GigaScience
4
,
7
.
Chen
,
W.
,
Burgess
,
S.
and
Hopkins
,
N.
(
2001
).
Analysis of the zebrafish smoothened mutant reveals conserved and divergent functions of hedgehog activity
.
Development
128
,
2385
-
2396
.
Cheng
,
X.
,
Collins
,
R. E.
and
Zhang
,
X.
(
2005
).
Structural and sequence motifs of protein (Histone) methylation enzymes
.
Annu. Rev. Biophys. Biomol. Struct.
34
,
267
-
294
.
Choi
,
H. M. T.
,
Calvert
,
C. R.
,
Husain
,
N.
,
Huss
,
D.
,
Barsi
,
J. C.
,
Deverman
,
B. E.
,
Hunter
,
R. C.
,
Kato
,
M.
,
Lee
,
S. M.
,
Abelin
,
A. C. T.
et al. 
(
2016
).
Mapping a multiplexed zoo of mRNA expression
.
Development
143
,
3632
-
3637
.
Choi
,
H. M. T.
,
Schwarzkopf
,
M.
,
Fornace
,
M. E.
,
Acharya
,
A.
,
Artavanis
,
G.
,
Stegmaier
,
J.
,
Cunha
,
A.
and
Pierce
,
N. A.
(
2018
).
Third-generation in situ hybridization chain reaction: multiplexed, quantitative, sensitive, versatile, robust
.
Development
145
,
dev165753
.
Corallo
,
D.
,
Schiavinato
,
A.
,
Trapani
,
V.
,
Moro
,
E.
,
Argenton
,
F.
and
Bonaldo
,
P.
(
2013
).
Emilin3 is required for notochord sheath integrity and interacts with Scube2 to regulate notochord-derived Hedgehog signals
.
Development
140
,
4594
-
4601
.
Correa-Cerro
,
L.
,
Garciaz-Cruz
,
D.
,
Diaz-Castanos
,
L.
,
Figuera
,
L.
and
Sanchez-Corona
,
J.
(
1996
).
Interstitial deletion 6q16.2q22.2 in a child with ectrodactyl
.
Ann. Genet.
39
,
105
-
109
.
Crackower
,
M.
,
Scherer
,
S.
,
Rommens
,
J.
,
Hui
,
C.
,
Poorkaj
,
P.
,
Soder
,
S.
,
Cobben
,
J.
,
Hudgins
,
L.
,
Evans
,
J.
and
Tsui
,
L.
(
1996
).
Characterization of the split hand/split foot malformation locus SHFM1 at 7q21.3-q22.1 and analysis of a candidate gene for its expression during limb development
.
Hum. Mol. Genet.
5
,
571
-
579
.
Crossley
,
P. H.
,
Minowada
,
G.
,
MacArthur
,
C. A.
and
Martin
,
G. R.
(
1996
).
Roles for FGF8 in the induction, initiation, and maintenance of chick limb development
.
Cell
84
,
127
-
136
.
Dale
,
R. M.
and
Topczewski
,
J.
(
2011
).
Identification of an evolutionarily conserved regulatory element of the zebrafish col2a1a gene
.
Dev. Biol.
357
,
518
-
531
.
Danecek
,
P.
,
Bonfield
,
J. K.
,
Liddle
,
J.
,
Marshall
,
J.
,
Ohan
,
V.
,
Pollard
,
M. O.
,
Whitwham
,
A.
,
Keane
,
T.
,
McCarthy
,
S. A.
,
Davies
,
R. M.
et al. 
(
2021
).
Twelve years of SAMtools and BCFtools
.
GigaScience
10
,
giab008
.
Dobin
,
A.
,
Davis
,
C. A.
,
Schlesinger
,
F.
,
Drenkow
,
J.
,
Zaleski
,
C.
,
Jha
,
S.
,
Batut
,
P.
,
Chaisson
,
M.
and
Gingeras
,
T.
(
2013
).
STAR: ultrafast universal RNA-seq aligner
.
Bioinformatics
29
,
15
-
21
.
Duijf
,
P. H. G.
,
van Bokhoven
,
H.
and
Brunner
,
H. G.
(
2003
).
Pathogenesis of split-hand/split-foot malformation
.
Hum. Mol. Genet.
12
,
R51
-
R60
.
Duran-Gonzalez
,
J.
,
Gutierrez-Angulo
,
M.
,
Garcia-Cruz
,
D.
,
de la Luz Ayala
,
M.
,
Padilla
,
M.
and
Davalos
,
I. P.
(
2007
).
A de novo interstitial 6q deletion in a boy with a split hand malformation
.
J. Appl. Genet.
48
,
405
-
407
.
Fulton
,
T.
,
Trivedi
,
V.
,
Attardi
,
A.
,
Anlas
,
K.
,
Dingare
,
C.
,
Arias
,
A. M.
and
Steventon
,
B.
(
2020
).
Axis specification in zebrafish is robust to cell mixing and reveals a regulation of pattern formation by morphogenesis
.
Curr. Biol.
30
,
2984
-
2994.e2983
.
Gegenbaur
,
C.
(
1878
).
Elements of Comparative Anatomy
.
Macmillan
.
Gene Ontology Consortium
. (
2021
).
The Gene Ontology resource: enriching a GOld mine
.
Nucleic Acids Res.
49
,
D325
-
D334
.
Gordon
,
L.
,
Joo
,
J. E.
,
Powell
,
J. E.
,
Ollikainen
,
M.
,
Novakovic
,
B.
,
Li
,
X.
,
Andronikos
,
R.
,
Cruickshank
,
M. N.
,
Conneely
,
K. N.
,
Smith
,
A. K.
et al. 
(
2012
).
Neonatal DNA methylation profile in human twins is specified by a complex interplay between intrauterine environmental and genetic factors, subject to tissue-specific influence
.
Genome Res.
22
,
1395
-
1406
.
Grandel
,
H.
and
Schulte-Merker
,
S.
(
1998
).
The development of the paired fins in the Zebrafish (Danio rerio)
.
Mech. Dev.
79
,
99
-
120
.
Grandel
,
H.
,
Lun
,
K.
,
Rauch
,
G.-J.
,
Rhinn
,
M.
,
Piotrowski
,
T.
,
Houart
,
C.
,
Sordino
,
P.
,
Küchler
,
A. M.
,
Schulte-Merker
,
S.
,
Geisler
,
R.
et al. 
(
2002
).
Retinoic acid signalling in the zebrafish embryo is necessary during pre-segmentation stages to pattern the anterior-posterior axis of the CNS and to induce a pectoral fin bud
.
Development
129
,
2851
-
2865
.
Gurrieri
,
F.
and
Everman
,
D. B.
(
2013
).
Clinical, genetic, and molecular aspects of split-hand/foot malformation: an update
.
Am. J. Med. Genet. A
161
,
2860
-
2872
.
Gurrieri
,
F.
,
Cammarata
,
M.
,
Avarello
,
R. M.
,
Genuardi
,
M.
,
Pomponi
,
M. G.
,
Neri
,
G.
and
Giuffrè
,
L.
(
1995
).
Ulnar ray defect in an infant with a 6q21;7q31.2 translocation: further evidence for the existence of a limb defect gene in 6q21
.
Am. J. Hum. Genet.
55
,
315
-
318
.
Gyory
,
I.
,
Wu
,
J.
,
Fejer
,
G.
,
Seto
,
E.
and
Wright
,
K. L.
(
2004
).
PRDI-BF1 recruits the histone H3 methyltransferase G9a in transcriptional silencing
.
Nat. Immunol.
5
,
299
-
308
.
Ha
,
A. S.
and
Riddle
,
R. D.
(
2003
).
cBlimp-1 expression in chick limb bud development
.
Gene Expr. Patterns
3
,
297
-
300
.
Heinz
,
S.
,
Benner
,
C.
,
Spann
,
N.
,
Bertolino
,
E.
,
Lin
,
Y. C.
,
Laslo
,
P.
,
Cheng
,
J. X.
,
Murre
,
C.
,
Singh
,
H.
and
Glass
,
C.
(
2010
).
Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities
.
Mol. Cell
38
,
576
-
589
.
Hernandez-Lagunas
,
L.
,
Choi
,
I. F.
,
Kaji
,
T.
,
Simpson
,
P.
,
Hershey
,
C.
,
Zhou
,
Y.
,
Zon
,
L.
,
Mercola
,
M.
and
Artinger
,
K. B.
(
2005
).
Zebrafish narrowminded disrupts the transcription factor prdm1 and is required for neural crest and sensory neuron specification
.
Dev. Biol.
278
,
347
-
357
.
Hernández-Vega
,
A.
and
Minguillón
,
C.
(
2011
).
The Prx1 limb enhancers: targeted gene expression in developing zebrafish pectoral fins
.
Dev. Dyn.
240
,
1977
-
1988
.
Heude
,
E.
,
Shaikho
,
S.
and
Ekker
,
M.
(
2014
).
The dlx5a/dlx6a genes play essential roles in the early development of zebrafish median fin and pectoral structures
.
PLoS ONE
9
,
e98505
.
Hohenauer
,
T.
and
Moore
,
A. W.
(
2012
).
The Prdm family: expanding roles in stem cells and development
.
Development
139
,
2267
-
2282
.
Hopkin
,
R. J.
,
Schorry
,
E.
,
Bofinger
,
M.
,
Milatovich
,
A.
,
Stern
,
H. J.
,
Jayne
,
C.
and
Saal
,
H. M.
(
1997
).
New insights into the phenotypes of 6q deletions
.
Am. J. Hum. Genet.
70
,
377
-
386
.
Janeway
,
C.
,
Travers
,
P.
and
Walport
,
M.
(
2001
).
Immunobiology: The Immune System in Health and Disease
, 5th edn.
Garland Science
. .
Kague
,
E.
,
Gallagher
,
M.
,
Burke
,
S.
,
Parsons
,
M.
,
Franz-Odendaal
,
T.
and
Fisher
,
S.
(
2012
).
Skeletogenic fate of zebrafish cranial and trunk neural crest
.
PLoS ONE
7
,
1
-
13
.
Kallies
,
A.
,
Hasbold
,
J.
,
Tarlinton
,
D. M.
,
Dietrich
,
W.
,
Corcoran
,
L. M.
,
Hodgkin
,
P. D.
and
Nutt
,
S. L.
(
2004
).
Plasma cell ontogeny defined by quantitative changes in blimp-1 expression
.
J. Exp. Med.
200
,
967
-
977
.
Kantaputra
,
P. N.
and
Carlson
,
B. M.
(
2019
).
Genetic regulatory pathways of split-hand/foot malformation
.
Clin. Genet.
95
,
132
-
139
.
Karczewski
,
K. J.
,
Francioli
,
L. C.
,
Tiao
,
G.
,
Cummings
,
B. B.
,
Alfoldi
,
J.
,
Wang
,
Q.
,
Collins
,
R. L.
,
Laricchia
,
K. M.
,
Ganna
,
A.
,
Birnbaum
,
D. P.
et al. 
(
2020
).
The mutational constraint spectrum quantified from variation in 141,456 humans
.
Nature
581
,
434
-
443
.
Kawakami
,
K.
,
Takeda
,
H.
,
Kawakami
,
N.
,
Kobayashi
,
M.
,
Matsuda
,
N.
and
Mishina
,
M.
(
2004a
).
A transposon-mediated gene trap approach identifies developmentally regulated genes in zebrafish
.
Dev. Cell
7
,
133
-
144
.
Kawakami
,
Y.
,
Esteban
,
C. R.
,
Matsui
,
T.
,
Rodríguez-León
,
J.
,
Kato
,
S.
and
Izpisúa Belmonte
,
J. C.
(
2004b
).
Sp8 and Sp9, two closely related buttonhead-like transcription factors, regulate Fgf8 expression and limb outgrowth in vertebrate embryos
.
Development
131
,
4763
-
4774
.
Keller
,
A. D.
and
Maniatis
,
T.
(
1992
).
Only two of the five zinc fingers of the eukaryotic transcriptional repressor PRDI-BF1 are required for sequence-specific DNA binding
.
Mol. Cell. Biol.
12
,
1940
-
1949
.
Kimmel
,
C. B.
,
Ballard
,
W. W.
,
Kimmel
,
S. R.
,
Ullmann
,
B.
and
Schilling
,
T. F.
(
1995
).
Stages of embryonic development of the zebrafish
.
Dev. Dyn.
203
,
253
-
310
.
Kouwenhoven
,
E. N.
,
van Heeringen
,
S. J.
,
Tena
,
J. J.
,
Oti
,
M.
,
Dutilh
,
B. E.
,
Alonso
,
M. E.
,
de la Calle-Mustienes
,
E.
,
Smeenk
,
L.
,
Rinne
,
T.
,
Parsaulian
,
L.
et al. 
(
2010
).
Genome-wide profiling of p63 DNA-binding sites identifies an element that regulates gene expression during limb development in the 7q21 SHFM1 locus
.
PLoS Genet.
6
,
e1001065
.
Kumar
,
R. A.
,
Everman
,
D. B.
,
Morgan
,
C. T.
,
Slavotinek
,
A.
,
Schwartz
,
C. E.
and
Simpson
,
E.
(
2007
).
Absence of mutations in NR2E1 and SNX3 in five patients with MMEP (microcephaly, microphthalmia, ectrodactyly, and prognathism) and related phenotypes
.
BMC Med. Genet.
8
,
48
.
Kwan
,
K. M.
,
Fujimoto
,
E.
,
Grabher
,
C.
,
Mangum
,
B. D.
,
Hardy
,
M. E.
,
Campbell
,
D. S.
,
Parant
,
J. M.
,
Yost
,
H. J.
,
Kanki
,
J. P.
and
Chien
,
C.-B.
(
2007
).
The Tol2kit: a multisite gateway-based construction kit for Tol2 transposon transgenesis constructs
.
Dev. Dyn.
236
,
3088
-
3099
.
Langmead
,
B.
and
Salzberg
,
S. L.
(
2012
).
Fast gapped-read alignment with Bowtie 2
.
Nat. Methods
9
,
357
-
359
.
Lee
,
B. C.
and
Roy
,
S.
(
2006
).
Blimp-1 is an essential component of the genetic program controlling development of the pectoral limb bud
.
Dev. Biol.
300
,
623
-
634
.
Leerberg
,
D. M.
,
Hopton
,
R. E.
and
Draper
,
B. W.
(
2019
).
Fibroblast growth factor receptors function redundantly during zebrafish embryonic development
.
Genetics
212
,
1301
-
1319
.
Lewandoski
,
M.
,
Sun
,
X.
and
Martin
,
G. R.
(
2000
).
Fgf8 signalling from the AER is essential for normal limb development
.
Nat. Genet.
26
,
460
-
463
.
Liao
,
Y.
,
Smyth
,
G. K.
and
Shi
,
W.
(
2014
).
featureCounts: an efficient general purpose program for assigning sequence reads to genomic features
.
Bioinformatics
30
,
923
-
930
.
Liu
,
X.
,
Wu
,
C.
,
Li
,
C.
and
Boerwinkle
,
E.
(
2016
).
dbNSFP v3.0: a one-stop database of functional predictions and annotations for human nonsynonymous and splice-site SNVs
.
Hum. Mutat.
37
,
235
-
241
.
Liu
,
X.
,
Li
,
C.
,
Mou
,
C.
,
Dong
,
Y.
and
Tu
,
Y.
(
2020
).
dbNSFP v4: a comprehensive database of transcript-specific functional predictions and annotations for human nonsynonymous and splice-site SNVs
.
Genome Med.
12
,
103
.
Lo Iacono
,
N.
,
Mantero
,
S.
,
Chiarelli
,
A.
,
Garcia
,
E.
,
Mills
,
A. A.
,
Morasso
,
M. I.
,
Costanzo
,
A.
,
Levi
,
G.
,
Guerrini
,
L.
and
Merlo
,
G. R.
(
2008
).
Regulation of Dlx5 and Dlx6 gene expression by p63 is involved in EEC and SHFM congenital limb defects
.
Development
135
,
1377
-
1388
.
Love
,
M. I.
,
Huber
,
W.
and
Anders
,
S.
(
2014
).
Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2
.
Genome Biol.
15
,
550
.
Manichaikul
,
A.
,
Mychaleckyj
,
J. C.
,
Rich
,
S. S.
,
Daly
,
K.
,
Sale
,
M.
and
Chen
,
W.-M.
(
2010
).
Robust relationship inference in genome-wide association studies
.
Bioinformatics
26
,
2867
-
2873
.
Martin
,
M.
(
2011
).
Cutadapt removes adapter sequences from high-throughput sequencing reads [next generation sequencing; small RNA; microRNA; adapter removal]
.
EMBnet. J.
17
,
3
.
Martin
,
C.
and
Zhang
,
Y.
(
2005
).
The diverse functions of histone lysine methylation
.
Nat. Rev. Mol. Cell Biol.
6
,
838
-
849
.
Matsuoka
,
T.
,
Ahlberg
,
P. E.
,
Kessaris
,
N.
,
Iannarelli
,
P.
,
Dennehy
,
U.
,
Richardson
,
W. D.
,
McMahon
,
A. P.
and
Koentges
,
G.
(
2005
).
Neural crest origins of the neck and shoulder
.
Nature
436
,
347
-
355
.
McLaren
,
W.
,
Gil
,
L.
,
Hunt
,
S. E.
,
Riat
,
H. S.
,
Ritchie
,
G. R. S.
,
Thormann
,
A.
,
Flicek
,
P.
and
Cunningham
,
F.
(
2016
).
The ensembl variant effect predictor
.
Genome Biol.
17
,
122
.
Meers
,
M. P.
,
Bryson
,
T. D.
,
Henikoff
,
J. G.
and
Henikoff
,
S.
(
2019
).
Improved CUT&RUN chromatin profiling tools
.
eLife
8
,
e46314
.
Mercader
,
N.
,
Fischer
,
S.
and
Neumann
,
C. J.
(
2006
).
Prdm1 acts downstream of a sequential RA, Wnt and Fgf signaling cascade during zebrafish forelimb induction
.
Development
133
,
2805
-
2815
.
Mi
,
H.
,
Muruganujan
,
A.
,
Ebert
,
D.
,
Huang
,
X.
and
Thomas
,
P. D.
(
2019
).
PANTHER version 14: more genomes, a new PANTHER GO — slim and improvements in enrichment analysis tools
.
Nucleic Acids Res.
47
,
D419
-
D426
.
Milanetto
,
M.
,
Tiso
,
N.
,
Braghetta
,
P.
,
Volpin
,
D.
,
Argenton
,
F.
and
Bonaldo
,
P.
(
2007
).
Emilin genes are duplicated and dynamically expressed during zebrafish embryonic development
.
Dev. Dyn.
237
,
222
-
232
.
Mills
,
A. A.
,
Zheng
,
B.
,
Wang
,
X.-J.
,
Vogel
,
H.
,
Roop
,
D. R.
and
Bradley
,
A.
(
1999
).
p63 is a p53 homologue required for limb and epidermal morphogenesis
.
Nature
398
,
708
-
713
.
Mongera
,
A.
,
Singh
,
A. P.
,
Levesque
,
M. P.
,
Chen
,
Y.-Y.
,
Konstantinidis
,
P.
and
Nüsslein-Volhard
,
C.
(
2013
).
Genetic lineage labeling in zebrafish uncovers novel neural crest contributions to the head, including gill pillar cells
.
Development
140
,
916
-
925
.
Nakamura
,
T.
,
Gehrke
,
A. R.
,
Lemberg
,
J.
,
Szymaszek
,
J.
and
Shubin
,
N. H.
(
2016
).
Digits and fin rays share common developmental histories
.
Nature
537
,
225
-
228
.
Neumann
,
C. J.
,
Grandel
,
H.
,
Gaffield
,
W.
,
Schulte-Merker
,
S.
and
Nüsslein-Volhard
,
C.
(
1999
).
Transient establishment of anteroposterior polarity in the zebrafish pectoral fin bud in the absence of sonic hedgehog activity
.
Development
126
,
4817
-
4826
.
Ng
,
J. K.
,
Kawakami
,
Y.
,
Büscher
,
D.
,
Raya
,
A.
,
Itoh
,
T.
,
Koth
,
C. M.
,
Rodriguez Esteban
,
C.
,
Rodriguez-Leon
,
J.
,
Garrity
,
D. M.
,
Fishman
,
M. C.
et al. 
(
2002
).
The limb identity gene Tbx5 promotes limb initiation by interacting with Wnt2b and Fgf10
.
Development
129
,
5161
-
5170
.
Ohuchi
,
H.
,
Nakagawa
,
T.
,
Yamamoto
,
A.
,
Araga
,
A.
,
Ohata
,
T.
,
Ishimaru
,
Y.
,
Yoshioka
,
H.
,
Kuwana
,
T.
,
Nohno
,
T.
,
Yamasaki
,
M.
et al. 
(
1997
).
The mesenchymal factor, FGF10, initiates and maintains the outgrowth of the chick limb bud through interaction with FGF8, an apical ectodermal factor
.
Development
124
,
2235
-
2244
.
Paila
,
U.
,
Chapman
,
B. A.
,
Kirchner
,
R.
and
Quinlan
,
A. R.
(
2013
).
GEMINI: integrative exploration of genetic variation and genome annotations
.
PLoS Comput. Biol.
9
,
e1003153
.
Pandya
,
A.
,
Braverman
,
N.
,
Pyeritz
,
R. E.
,
Ying
,
K.-L.
,
Kline
,
A. D.
and
Falk
,
R. E.
(
1995
).
Interstitial deletion of the long arm of chromosome 6 associated with unusual limb anomalies: report of two new patients and review of the literature
.
Am. J. Hum. Genet.
59
,
38
-
43
.
Powell
,
D. R.
,
Hernandez-Lagunas
,
L.
,
LaMonica
,
K.
and
Artinger
,
K. B.
(
2013
).
Prdm1a directly activates foxd3 and tfap2a during zebrafish neural crest specification
.
Development
140
,
3445
-
3455
.
Prajapati
,
R. S.
,
Hintze
,
M.
and
Streit
,
A.
(
2019
).
PRDM1 controls the sequential activation of neural, neural crest and sensory progenitor determinants
.
Development
146
,
dev181107
.
Ramírez
,
F.
,
Dündar
,
F.
,
Diehl
,
S.
,
Grüning
,
B. A.
and
Manke
,
T.
(
2014
).
deepTools: a flexible platform for exploring deep-sequencing data
.
Nucleic Acids Res.
42
,
W187
-
W191
.
Ren
,
B.
,
Chee
,
K. J.
,
Kim
,
T. H.
and
Maniatis
,
T.
(
1999
).
PRDI-BF1/Blimp-1 repression is mediated by corepressors of the Groucho family of proteins
.
Genes Dev.
13
,
125
-
137
.
Riddle
,
R. D.
,
Johnson
,
R. L.
,
Laufer
,
E.
and
Tabin
,
C.
(
1993
).
Sonic hedgehog mediates the polarizing activity of the ZPA
.
Cell
75
,
1401
-
1416
.
Robertson
,
E. J.
,
Charatsi
,
I.
,
Joyner
,
C. J.
,
Koonce
,
C. H.
,
Morgan
,
M.
,
Islam
,
A.
,
Paterson
,
C.
,
Lejsek
,
E.
,
Arnold
,
S. J.
,
Kallies
,
A.
et al. 
(
2007
).
Blimp1 regulates development of the posterior forelimb, caudal pharyngeal arches, heart and sensory vibrissae in mice
.
Development
134
,
4335
-
4345
.
Robinson
,
J. T.
,
Thorvaldsdóttir
,
H.
,
Winckler
,
W.
,
Guttman
,
M.
,
Lander
,
E. S.
,
Getz
,
G.
and
Mesirov
,
J. P.
(
2011
).
Integrative genomics viewer
.
Nat. Biotechnol.
29
,
24
-
26
.
Robledo
,
R. F.
,
Rajan
,
L.
,
Li
,
X.
and
Lufkin
,
T.
(
2002
).
The Dlx5 and Dlx6 homeobox genes are essential for craniofacial, axial, and appendicular skeletal development
.
Genes Dev.
16
,
1089
-
1101
.
Rossi
,
C. C.
,
Kaji
,
T.
and
Artinger
,
K. B.
(
2009
).
Transcriptional control of Rohon-Beard sensory neuron development at the neural plate border
.
Dev. Dyn.
238
,
931
-
943
.
Roy
,
S.
and
Ng
,
T.
(
2004
).
Blimp-1 specifies neural crest and sensory neuron progenitors in the zebrafish embryo
.
Curr. Biol.
14
,
1772
-
1777
.
Roy
,
S.
,
Wolff
,
C.
and
Ingham
,
P. W.
(
2001
).
The u-boot mutation identifies a Hedgehog-regulated myogenic switch for fiber-type diversification in the zebrafish embryo
.
Genes Dev.
15
,
1563
-
1576
.
Rupp
,
R. A.
,
Snider
,
L.
and
Weintraub
,
H.
(
1994
).
Xenopus embryos regulate the nuclear localization of XMyoD
.
Genes Dev.
8
,
1311
-
1323
.
Saunders
,
J. W.
and
Gasseling
,
M. T.
(
1968
). Ectodermal-mesenchymal interactions in the origin of limb symmetry. In:
Epithelial-Mesenchymal Interactions
(ed.
R.
Fleischmajer
and
R. F.
Billingham
), pp.
78
-
97
. Baltimore: Williams & Wilkins.
Scherer
,
S.
,
Poorkaj
,
P.
,
Massa
,
H.
,
Soder
,
S.
,
Allen
,
T.
,
Nunes
,
M.
,
Geshurl
,
D.
,
Wong
,
E.
,
Bellonl
,
E.
,
Little
,
S.
et al. 
(
1994
).
Physical mapping of the split hand/split foot locus on chromosome 7 and implication in syndromic ectrodactyly
.
Hum. Mol. Genet.
3
,
1345
-
1354
.
Schwarz
,
J. M.
,
Rödelsperger
,
C.
,
Schuelke
,
M.
and
Seelow
,
D.
(
2010
).
MutationTaster evaluates disease-causing potential of sequence alterations
.
Nat. Methods
7
,
575
-
576
.
Shull
,
L. C.
,
Lencer
,
E. S.
,
Kim
,
H. M.
,
Goyama
,
S.
,
Kurokawa
,
M.
,
Costello
,
J. C.
,
Jones
,
K.
and
Artinger
,
K. B.
(
2022
).
PRDM paralogs antagonistically balance Wnt/β-catenin activity during craniofacial chondrocyte differentiation
.
Development
149
,
dev200082
.
Skene
,
P. J.
and
Henikoff
,
S.
(
2017
).
An efficient targeted nuclease strategy for high-resolution mapping of DNA binding sites
.
eLife
6
,
e21856
.
Sleight
,
V. A.
and
Gillis
,
J. A.
(
2020
).
Embryonic origin and serial homology of gill arches and paired fins in the skate, Leucoraja erinacea
.
eLife
9
,
e60635
.
Sordino
,
P.
,
van der Hoeven
,
F.
and
Duboule
,
D.
(
1995
).
Hox gene expression in teleost fins and the origin of vertebrate digits
.
Nature
375
,
678
-
681
.
Sordino
,
P.
,
Duboule
,
D.
and
Kondo
,
T.
(
1996
).
Zebrafish Hoxa and Evx-2 genes: cloning, developmental expression and implications for the functional evolution of posterior Hox genes
.
Mech. Dev.
59
,
165
-
175
.
Sowinska-Seidler
,
A.
,
Socha
,
M.
and
Jamsheer
,
A.
(
2014
).
Split-hand/foot malformation - molecular cause and implications in genetic counseling
.
J. Appl. Genet.
55
,
105
-
115
.
Sternberg
,
S. R.
(
1983
).
Biomedical image processing
.
Computer
16
,
22
-
34
.
Su
,
S.-T.
,
Ying
,
H.-Y.
,
Chiu
,
Y.-K.
,
Lin
,
F.-R.
,
Chen
,
M.-Y.
and
Lin
,
K.-I.
(
2009
).
Involvement of histone demethylase LSD1 in Blimp-1-mediated gene repression during plasma cell differentiation
.
Mol. Cell. Biol.
29
,
1421
-
1431
.
Talbot
,
J. C.
,
Johnson
,
S. L.
and
Kimmel
,
C. B.
(
2010
).
hand2 and Dlx genes specify dorsal, intermediate and ventral domains within zebrafish pharyngeal arches
.
Development
137
,
2507
-
2517
.
The 1000 Genomes Project Consortium
(
2015
).
A global reference for human genetic variation
.
Nature
526
,
68
-
74
.
Truong
,
B. T.
and
Artinger
,
K. B.
(
2021
).
The power of zebrafish models for understanding the co-occurrence of craniofacial and limb disorders
.
Genesis
59
,
e23407
.
Tsukahara
,
M.
,
Yoneda
,
J.
,
Azuma
,
R.
,
Nakashima
,
K.
,
Kito
,
N.
,
Ouchi
,
K.
and
Kanehara
,
Y.
(
1997
).
Interstitial deletion of 6q21–q23 associated with split hand
.
Am. J. Hum. Genet.
69
,
268
-
270
.
Turner
,
D. L.
and
Weintraub
,
H.
(
1994
).
Expression of achaete-scute homolog 3 in Xenopus embryos converts ectodermal cells to a neural fate
.
Genes Dev.
8
,
1434
-
1447
.
Ullah
,
A.
,
Hammid
,
A.
,
Umair
,
M.
and
Ahmad
,
W.
(
2017
).
A novel heterozygous intragenic sequence variant in DLX6 probably underlies first case of autosomal dominant split-hand/foot malformation type 1
.
Mol. Syndromol.
8
,
79
-
84
.
Umair
,
M.
and
Hayat
,
A.
(
2020
).
Nonsyndromic split-hand/foot malformation: recent classification
.
Mol. Syndromol.
10
,
243
-
254
.
Viljoen
,
D.
and
Smart
,
R.
(
1993
).
Split-foot anomaly, microphthalmia, cleft-lip and cleft-palate, and mental retardation associated with a chromosome 6;13 translocation
.
Clin. Dysmorphol.
2
,
274
-
277
.
Vincent
,
S. D.
,
Dunn
,
N. R.
,
Sciammas
,
R.
,
Shapiro-Shalef
,
M.
,
Davis
,
M. M.
,
Calame
,
K.
,
Bikoff
,
E. K.
and
Robertson
,
E. J.
(
2005
).
The zinc finger transcriptional repressor Blimp1/Prdm1 is dispensable for early axis formation but is required for specification of primordial germ cells in the mouse
.
Development
132
,
1315
-
1325
.
von Hofsten
,
J.
,
Elworthy
,
S.
,
Gilchrist
,
M. J.
,
Smith
,
J. C.
,
Wardle
,
F. C.
and
Ingham
,
P. W.
(
2008
).
Prdm1- and Sox6-mediated transcriptional repression specifies muscle fibre type in the zebrafish embryo
.
EMBO Rep.
9
,
683
-
689
.
Walker
,
M.
and
Kimmel
,
C.
(
2007
).
A two-color acid-free cartilage and bone stain for zebrafish larvae
.
Biotech. Histochem.
82
,
23
-
28
.
Wang
,
X.
,
Xin
,
Q.
,
Li
,
L.
,
Li
,
J.
,
Zhang
,
C.
,
Qiu
,
R.
,
Qian
,
C.
,
Zhao
,
H.
,
Liu
,
Y.
,
Shan
,
S.
et al. 
(
2014
).
Exome sequencing reveals a heterozygous DLX5 mutation in a Chinese family with autosomal-dominant split-hand/foot malformation
.
Eur. J. Hum. Genet.
22
,
1105
-
1110
.
Westerfield
,
M.
(
2000
).
The Zebrafish Book. A Guide for the Laboratory Use of Zebrafish (Danio rerio)
, 4th edn.
[Online Book]
.
University of Oregon Press
. .
Wilcox
,
W. R.
,
Coulter
,
C. P.
and
Schmitz
,
M. L.
(
2015
).
Congenital limb deficiency disorders
.
Clin. Perinatol.
42
,
281
-
300
,
viii
.
Wilm
,
T. P.
and
Solnica-Krezel
,
L.
(
2005
).
Essential roles of a zebrafish prdm1/blimp1 homolog in embryo patterning and organogenesis
.
Development
132
,
393
-
404
.
Wu
,
T.
,
Hu
,
E.
,
Xu
,
S.
,
Chen
,
M.
,
Guo
,
P.
,
Dai
,
Z.
,
Feng
,
T.
,
Zhou
,
L.
,
Tang
,
W.
,
Zhan
,
L.
et al. 
(
2021
).
clusterProfiler 4.0: a universal enrichment tool for interpreting omics data
.
The Innovation
2
,
100141
.
Yan
,
Y.-L.
,
Hatta
,
K.
,
Riggleman
,
B.
and
Postlethwait
,
J. H.
(
1995
).
Expression of a type II collagen gene in the zebrafish embryonic axis
.
Dev. Dyn.
203
,
363
-
376
.
Yang
,
A.
,
Schweitzer
,
R.
,
Sun
,
D.
,
Kaghad
,
M.
,
Walker
,
N.
,
Bronson
,
R. T.
,
Tabin
,
C.
,
Sharpe
,
A.
,
Caput
,
D.
,
Crum
,
C.
et al. 
(
1999
).
p63 is essential for regenerative proliferation in limb, craniofacial, and epithelial development
.
Nature
398
,
714
-
718
.
Yano
,
T.
and
Tamura
,
K.
(
2013
).
The making of differences between fins and limbs
.
J. Anat.
222
,
100
-
113
.
Ye
,
Z.
,
Braden
,
C. R.
,
Wills
,
A.
and
Kimelman
,
D.
(
2021
).
Identification of in vivo Hox13-binding sites reveals an essential locus controlling zebrafish brachyury expression
.
Development
148
,
dev199408
.
Yelon
,
D.
,
Ticho
,
B.
,
Halpern
,
M. E.
,
Ruvinsky
,
I.
,
Ho
,
R. K.
,
Silver
,
L. M.
and
Stainier
,
D. Y. R.
(
2000
).
The bHLH transcription factor hand2 plays parallel roles in zebrafish heart and pectoral fin development
.
Development
127
,
2573
-
2582
.
Yu
,
J.
,
Angelin-Duclos
,
C.
,
Greenwood
,
J.
,
Liao
,
J.
and
Calame
,
K.
(
2000
).
Transcriptional repression by Blimp-1 (PRDI-BF1) involves recruitment of histone deacetylase
.
Mol. Cell. Biol.
20
,
2592
-
2603
.
Yu
,
G.
,
Wang
,
L.-G.
and
He
,
Q.-Y.
(
2015
).
ChIPseeker: an R/Bioconductor package for ChIP peak annotation, comparison, and visualization
.
Bioinformatics
31
,
2382
-
2383
.
Zhang
,
Y.
,
Liu
,
T.
,
Meyer
,
C. A.
,
Eeckhoute
,
J.
,
Johnson
,
D. S.
,
Bernstein
,
B. E.
,
Nusbaum
,
C.
,
Myers
,
R. M.
,
Brown
,
M.
,
Li
,
W.
et al. 
(
2008
).
Model-based analysis of ChIP-Seq (MACS)
.
Genome Biol.
9
,
R137
.

Competing interests

The authors declare no competing or financial interests.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution and reproduction in any medium provided that the original work is properly attributed.

Supplementary information