Abstract
Rapid advances and cost erosion in exome and genome analysis of patients with both rare and common genetic disorders has accelerated gene discovery and illuminated fundamental biological mechanisms. The thrill of discovery has been accompanied, however, with the sobering appreciation that human genomes are burdened with a large number of rare and ultra rare variants, thereby posing a significant challenge in dissecting both the effect of such alleles on protein function and the biological relevance of these events to patient pathology. In an effort to develop model systems that are able to generate surrogates of human pathologies, a powerful suite of tools have been developed in zebrafish, taking advantage of the relatively small (compared to invertebrate models) evolutionary distance of that genome to humans, the orthology of several organs and signaling processes, and the suitability of this organism for medium and high throughput phenotypic screening. Here we will review the use of this model organism in dissecting human genetic disorders; we will highlight how diverse strategies have informed disease causality and genetic architecture; and we will discuss relative strengths and limitations of these approaches in the context of medical genome sequencing.
Introduction: Challenges and Opportunities from Human Genetics
Major inflections in genomic advances have always been accompanied by accelerated discovery of lesions associated with human pathologies. The development of the first karyotype led rapidly to the discovery of syndromes of polyploidy [1], while the then nascent technologies of genome mapping, cloning and sequencing yielded early insights into rare disease pathogenesis [2]. As the field progressed, molecular cytogenetics at the sub-Mb and ultimately kb-level resolution revealed the high contribution of copy number variants to both rare and common human genetic disorders [3], while, most recently, whole exome and whole genome sequencing (WES/WGS) have hyper-accelerated disease gene discovery both in historical cohorts and in the real-time clinical setting [4].
Amid the euphoria of discovery and the acutely increased expectation from patients and their physicians that the application of genomics can accelerate diagnosis and focus treatment options, the sobering realization has also emerged that each individual human genome is burdened with a large number of rare and ultra rare alleles. Considering bona fide pathogenic mutations alone in the average human exome, studies have reported a median of 50–150 nonsense mutations, several in homozygosity, while the abundance of unique single nucleotide variants (SNVs) can be in in the low to mid 100s [5]. Importantly, the number of rare and ultra-rare SNVs has continued to increase proportionately to the number of available exomes and genomes [6], indicating that we are unlikely to reach saturation of such alleles soon. These observations have generated a significant interpretive problem for disease gene discovery and for clinical genomics, as population-based arguments alone have been unable to dissect the contribution of the majority of these alleles to clinical phenotypes. Computational algorithms that take into consideration a variety of evolutionary, structural and biophysical properties of proteins have been of some assistance; however, their predictive ability (estimated in the 70 – 80% range [7]) has remained somewhat limited, mandating that definitive assessment of pathogenicity must be carried out through other methods.
Animal studies combine the identification of candidate alleles for human diseases with mutant organisms that recapitulate the human mutation or loss of gene function, and have improved our understanding of the causal link between genetic mutation and phenotypic trait [8]. Numerous animal models have been developed to study both monogenic and complex disease. Each model system has its advantages and limitations, such as genetic and anatomic homology to humans, the size of the genetic toolkit, generation time, and cost. Here we will focus on the application of zebrafish in modeling human genetic disease, an organism that has gained utility by bridging the gap between the high throughput abilities of invertebrates and the orthology of structure of mammals (Tables 1, 2). Although not a panacea, the implementation of zebrafish complementation studies [9] - suppression of the orthologous zebrafish gene and rescue with either a mutant or wild-type human mRNA to determine pathogenicity - in human and medical genomics has facilitated disease gene discovery in both monogenic and complex traits, and has also found application in modeling more intricate (and challenging) genetic lesions that include CNVs and epistatic interactions. We will review the tools available, discuss their possible uses and limitations and place the current vector of development of this model organism in the context of the ever- expanding generation of patient genomic data and the need for their accurate interpretation.
Table 1.
General attributes similarities of laboratory organisms used to model human genetic disease
C. elegans | D. melanogaster | D. rerio | M. musculus | |
---|---|---|---|---|
Percent identity with H. sapiens | 43% | 61% | 70% | 80% |
Genome size | 9.7 x 107 bp | 1.3 x 108 bp | 1.4 x 109 bp | 2.5 x 109 bp |
Exome size | 28.1 Mb | 30.9 Mb | 96 Mb | 49.6 Mb |
Practical attributes | ||||
Husbandry demands | $ | $ | $ | $$$ |
Cost per animal | $ | $ | $ | $$$ |
Characterized inbred strains | + | + | + | ++++ |
Outbred laboratory strains | + | + | +++ | ++ |
Germline/Embryonic cryopreservation | Yes | No | Yes | Yes |
Lifespan | 2 weeks | 0.3 years | 2–3 years | 1.3–3 years |
Generation interval | 5.5 days | 2 weeks | 3 months | 6–8 weeks |
Number of offspring | 300 larva | 10–20 eggs | >200 embryos/clutch | 10–12 pups/litter |
Embryonic development | ex vivo | ex vivo | ex vivo | in utero |
Molecular biology tools | ||||
Transgenesis * | ++++ | +++ | ++++ | +++ |
Gene targeting * | ++++ | +++ | + | ++++ |
Conditional gene targeting | + | ++ | + | ++++ |
Transient in vivo assays * | +++ | ++ | ++++ | + |
Allelic series from TILLING * | +++ | +++ | ++++ | ++ |
Affordability of large scale screens ** | ++++ | ++++ | +++ | + |
Cell biology tools | ||||
Cell lines and tissue culture | + | ++ | + | ++++ |
Antibody reagents | + | ++ | + | ++++ |
In situ probes | + | +++ | ++++ | +++ |
Disease process | ||||
Birth defects | + | ++ | ++++ | ++++ |
Adult-onset | ++ | + | + | ++++ |
Behavioral | ++ | ++ | ++ | ++ |
Aging | +++ | ++ | ++ | ++ |
Metabolic | ++ | ++ | +++ | +++ |
Table adapted from Lieschke and Currie 2007. Nat Rev Genet. 8: 353–367.
reverse genetics;
forward genetic
Table 2.
Anatomical comparisons between zebrafish and humans
Anatomy | Key similarities | Key differences |
---|---|---|
Embryology |
|
|
Skeletal system |
|
|
Muscle |
|
|
Nervous system and behavior |
|
|
Hematopoietic and lymphoid/immune systems |
|
|
Cardiovascular system |
|
|
Respiratory system |
|
|
Gastrointestinal system |
|
|
Renal and urinary systems |
|
|
Reproductive system |
|
|
Endocrine system |
|
|
Skin and appendages |
|
|
Table adapted from Lieschke and Currie 2007. Nat Rev Genet. 8: 353–367.
Animal models of human genetic disease
A deep understanding of the genetic architecture of human disease, underlying cellular and molecular mechanisms, and the development of therapeutic paradigms is dependent on model organisms that can robustly capture the pathology under investigation. Mammalian models such as the mouse (Mus musculus), have historically been attractive platforms by virtue of a high level of genomic sequence homology to humans (>80%) [10], highly conserved anatomical and physiological features, and a diverse repertoire of gene targeting strategies to recapitulate human disease phenotypes (for scholarly reviews of this topic see [11, 12]). However, in the context of human genomics, time and cost now present significant drawbacks. By contrast, invertebrate models such as the nematode worm (Caenorhabditis elegans), or the fruit fly (Drosophila melanogaster) offer inexpensive alternatives, and have proven especially powerful for studying orthologous genes of interest through the use of sophisticated gene manipulation strategies (RNAi, transposable insertion elements etc; refs [13, 14] for reviews). Nonetheless, the reduced cost and increased experimental tractability of these models are accompanied with a greater disparity in cellular processes and structures in comparison to humans, in large part due to a decreased percentage of genes shared between species (43% and 61% for worm and fly versus human, respectively [15]). An intermediate model, the zebrafish (Danio rerio), has emerged as strong candidate to achieve the experimental tractability of its invertebrate counterparts, but with the genomic and physiological proximity of a vertebrate for the investigation of human genetic disease.
Zebrafish: an overview
The zebrafish is a tropical teleost that lives in the freshwaters of south-east Asia. In the late 1960s, George Streisinger transitioned this common aquarium species to a model for basic research of embryogenesis and organ development because of its “desirable attributes,” including a relatively short generation time of three to four months, the ability of mating pairs to generate several hundred embryos that develop rapidly and synchronously ex vivo, and the small size of adult fish (3 cm in length), making them easy to care for [16]. Moreover, embryos are transparent, allowing facile microscopic visualization in the first days of development, with major organ formation occurring 24 hours post-fertilization. Zebrafish have a diploid genome, but differ notably from the genomic structure of other vertebrates by the major teleost specific genome duplication that has resulted in sub-functionalization and neo-functionalization of genes [17–19]. Importantly, the biomedical research community now has a publicly available, extensively annotated version of the zebrafish genome at its disposal, of which 70% of genes have an identifiable human ortholog [20]. Additionally, a vast catalog of mutants, transgenic reporters, and gene-specific expression data has been generated from over two decades of dedicated Danio rerio use for “phenotype-driven” forward genetic screens and “gene-driven” reverse genetic approaches. These data are curated in ZFIN (the Zebrafish Model Organism Database), a community-wide resource warehousing genomic information, anatomical atlases, molecular tools and links to zebrafish publications (www.zfin.org and [21]).
Forward genetics: advances in vertebrate developmental biology
Initial forays in zebrafish research predated the precise knowledge of gene content or location within the zebrafish genome, and were not necessarily motivated by targeted questions of human pathology. Rather, most forward screens were conducted to understand vertebrate embryonic development by 1) introducing random mutations throughout the genome; 2) conducting an informative breeding scheme to generate progeny with homozygous recessive mutations; 3) evaluating animals for a measurable phenotypic readout; and 4) identifying the mutation and gene underscoring the phenotype. Used widely across multiple model organisms, the application of this traditionally laborious strategy in zebrafish has been reviewed extensively elsewhere [22].
The first zebrafish screens were reported in the 1980s and involved the application of gamma rays to induce recessive lethal mutations. However, this approach resulted in significant chromosomal breaks that rendered mapping to a single locus difficult [23, 24]. Alkylating agents, primarily N-ethyl-N-nitrosourea (ENU), replaced gamma rays as an effective mutagen and application resulted in discrete genomic mutagenesis in zebrafish germ cells that could be mapped to a single gene (approximately one mutant per genome evaluated) [25, 26]. This discovery led to large-scale efforts by labs in Tübingen and Boston to apply ENU screening to zebrafish. Within two years, their combined efforts led to the characterization of ~4,000 embryonic lethal phenotypes; these include gastrulation [27, 28]; somitogenesis [29]; brain [30–32]; cardiovascular [33]; and craniofacial development mutants [34–36].
Although forward genetic screens in zebrafish contributed significantly to the fundamental understanding of early embryonic development, the impact on such findings to inherited disease in humans has been sporadic. This modest connection can be attributed to four main reasons. First, such screens are unable to capture alleles that confer a dominant negative (antagonizes the wild-type protein function) or gain of function (mutation confers a protein function different from that of wild-type protein) effect. Second, phenotypes must have a measurable phenotypic readout in early embryonic or larval stages, decreasing the possibility of detecting adult-onset or degenerative phenotypes. Third, although such screens were able to uncover discrete gene functions, the odds of generating precisely the same allele by ENU as has been seen in a patient is remote. Finally, this approach is confounded further by the fact that the zebrafish genome underwent a teleost specific duplication [17–19]. Among the genes for which there is an identifiable human ortholog, 47% have a one-to-one orthologous relationship with a human counterpart, while the remainder of zebrafish genes have complicated one-to-many or many-to-one orthology in comparison to the human gene [20]. As a result, duplicated gene function may either be a) retained in both copies making them functionally redundant; b) lost in one of the two copies wherein it becomes a pseudogene; or c) a novel and divergent gene function is acquired by one of the two copies. Therefore, mutations in only one of two functionally redundant orthologs might not display a phenotype. Finally, Nonetheless, ENU mutants have been successful in drawing anatomical correlates for genes implicated in recessive human disorders that cause anatomical birth defects. For instance, the craniofacial mutant crusherm299 is caused by a nonsense mutation in sec23a [37]; at the same time as this discovery, SEC23A mutations in humans were shown to cause a clinically relevant craniofacial dysmorphology, cranio-lenticulo-sutural dysplasia, bolstering the evidence of causality in both species [38]. Importantly, the recent application of WGS [39], WES [40], and improved mapping strategies [41] to zebrafish ENU mutants has enabled the rapid and cost-effective identification of mutations, justifying the continued use of forward genetics to assist with assigning causality in human genetic disease.
Reverse genetics: from candidate causal gene to physiologically relevant animal model
Forward genetic screening involves the unbiased examination of phenotypes resulting from mutations in the zebrafish genome. However, the randomness of this approach is hampered by the inability to specifically target every coding gene and/or specific mutations implicated in human pathology. To circumvent this problem, the precise targeting of candidate genes and alleles can be achieved through several methods that have been developed over the past ~15 years.
First, transient gene manipulation can be achieved through the injection of either morpholino (MO) antisense oligonucleotides (suppression) or capped in vitro transcribed mRNA (ectopic expression) into zebrafish embryos. MOs are stable molecules that consist of a large, nonribose morpholine backbone with the four DNA bases pairing stably with mRNA at either the translation start site (to disrupt protein synthesis) or at intron-exon boundaries (to disrupt mRNA splicing) [42]. The use of MOs to confer effective gene knockdown was first shown in zebrafish in 2000; Nasevicius and Ekker recapitulated the developmental phenotypes of five different embryonic lethal mutants and also developed models of the human genetic disorders hepatoerythropoietic porphyria and holoprosencephaly through the suppression of urod and shh, respectively [43]. Since this report, MOs have been used broadly to study vertebrate development and disease; co-injection of MO and orthologous mRNA has been employed for the systematic functional testing of alleles identified in humans, offering a powerful approach for analysis of variant pathogenicity and direction of effect [44]. Still, this methodology does have notable drawbacks: 1) MO efficacy is limited to ~3–5 days [43], and similarly, the presence of mRNA is limited to the same embryonic timeframe; 2) with few exceptions [45], injected MOs and mRNAs do not confer spatial or temporal specific activity; and 3) MOs can give rise to spurious phenotypes resulting from off-target effects [46]. Even so, the use of this methodology within the appropriate developmental stage, and with the appropriate experimental controls (a) targeting with a splice-blocking MO to demonstrate incorrectly spliced RNA; b) specific rescue of MO phenotypes with orthologous wild-type mRNA; c) demonstration of a similar phenotype with multiple MOs targeting the same gene; or d) comparison with a mutant when possible, and if appropriate [46]) can allow for the correct interpretation of MO phenotypes to establish relevance to human genetic disease through the recapitulation of loss-of-function or dominant negative effects.
Second, it is possible to readily obtain germ line zebrafish mutants for a gene of interest; this avoids the phenotypic variability associated with MOs, and allows the observation of phenotypes beyond early larval stages. Targeting Induced Local Lesions in Genomes (TILLING) was the first reverse genetic approach to produce germ line mutations in a gene of interest. Similar to forward screens, TILLING involves ENU mutagenesis of adult male zebrafish and generation of F1 families. Sperm from F1 males is then cryopreserved while genomic lesions are screened in target genes, typically in early exons or near exonic regions encoding critical protein domains, through PCR amplicon screening [47]. The completion of the zebrafish genome coupled to next-generation sequencing has increased significantly the throughput of the screening aspect of this strategy. TILLING mutants have been identified for >38% of all known zebrafish protein coding genes (http://www.sanger.ac.uk/Projects/D_rerio/zmp and [48]); this corresponds to ~60% of orthologous genes associated with a human phenotype in the Online Mendelian Inheritance in Man (OMIM; http://www.omim.org) database. The ongoing TILLING efforts hope to generate a comprehensive resource of putative null or hypomorphic models of human genetic disease, however, it is critical to be cognizant of the possibility that ENU may introduce multiple lesions in the genome. Ideally, multiple mutants with different alleles in the same gene should be phenotypically characterized to ensure that the pathology is specific. The same guidelines are true for retrovirus [49] or transposon [50] insertional mutants used in similar reverse genetics approaches.
Both forward ENU screens and TILLING are laborious; alternative approaches have recently expanded the utility of the zebrafish by enabling precise and germ line transmittable gene targeting that does not require excessive downstream screening to identify mutations (see ref [14] in this issue for a detailed review of genome editing endonucleases). First, zinc finger nucleases (ZFN) utilize a zinc finger array to enable target sequence specificity (typically the early exon of a gene), and a FokI endonuclease to guide cleavage and subsequent repair at the target site [51]; this was first utilized to target the gol locus (mutation of which results in absence of pigment), ntl (a regulator of early embryogenesis), and kdr (vascular endothelial growth factor-2 receptor), as visible proof-of-principle phenotypes [52, 53]. Second, transcription activator-like TAL effector nucleases (TALEN) have similarly been optimized to achieve locus-specific genome editing and have been shown to achieve greater specificity of and alteration of target sequences than ZFNs [54]. A third, more recent advancement in zebrafish genome editing technology involves clustered, regularly interspaced, short palindromic repeats (CRISPR), bacterial type II systems that guide RNAs to direct site-specific DNA cleavage by the Cas9 endonuclease [55]. Each of ZFN, TALEN, and CRISPR technologies have expanded the molecular toolkit of the zebrafish (for comparisons, see Table 3), accelerating studies of vertebrate development and improving our understanding of analogous phenotypes to human disease. For example, CRISPR/Cas9 was used to edit the gata5 locus [56], and mutant embryos displayed a cardia bifida phenotype mimicking both the fautm236a zebrafish mutant [57], and also humans with congenital heart defects [58–60]. However, there is currently a relative paucity of reports in which human-driven WES/WGS studies have been followed with the generation of such stable mutants; this is largely due to the relative newness of the technology, and/or the amount of time and labor still required to generate and characterize mutants; we anticipate the landscape of the field to change rapidly in the coming months and years.
Table 3.
Comparison of different reverse genetics approaches.
Methodology | Class | Specificity (targeting sequence) |
Off target effects |
Technology adoption time and costs |
Ongoing time and cost |
Somatic DNA cutting efficiency |
Germline efficacy |
---|---|---|---|---|---|---|---|
Morpholinos | Morpholino oligo |
25 nt | Low | Low | Low | n/a | n/a |
ZFNs | Protein-DNA | 18+ bp (pair) | Low | High | Moderate | Low (~2%) | Low |
TALENs | Protein-DNA | 30+ bp (pair) | Ultra low | Moderate | Low | Moderate to high (~20– ~50%) |
Moderate to high |
CRISPR- Cas9 |
RNA-DNA | 12+ bp | Ultra low | Low | Low | Moderate to high (~30%– ~60%) |
Moderate to high |
Table adapted from Blackburn et al. 2013 Zebrafish. 10: 116–118.
Humanizing the zebrafish to study mutations detected in humans
Taken together, the zebrafish exemplifies a tractable and physiologically relevant tool to model genetic variation in humans. Each of the forward and reverse genetics tools have limitations, and in particular, place significant emphasis towards the study of loss-of-function effects of single genes potentially making them an overly simplistic model to investigate oligogenic or even complex traits. In a growing number of instances, however, it has been possible to balance experimental tractability, specificity, and cross-species phenotypic similarity to establish: 1) physiological relevance of a gene to a human clinical phenotype; 2) allele pathogenicity; and 3) direction of allele effect for a vast array of human genetic disorders with diverse models of inheritance, phenotypes, and ages of onset.
Recessive disease
Disorders that segregate under a recessive mode of inheritance, especially congenital or pediatric-onset disorders with an abnormality in a structure with an anatomical counterpart in the developing zebrafish, have achieved widespread use towards demonstrating physiological relevance. This often represents the extent of functional data presented in instances when the human mutations have an unambiguous loss-of-function effect on the protein (nonsense, frameshift, or splice-site). For instance, causal mutations identified in primary ciliary Dyskinesia (PCD) are almost exclusively null changes, and transient MO-based studies in zebrafish have shown that proteins of a priori unknown function including CCDC39, ARMC4, and ZMYND10 give rise to left-right asymmetry defects phenotypes found in humans [61–63]. In other recessive disorders, such as pontocerebellar hypoplasia (PCH), the zebrafish has assisted in establishing clinical relevance and allele pathogenicity (Figure 1A). Wan et al. identified nonsynonymous changes in EXOSC3, encoding exosome component 3, following WES of four affected siblings; MO-induced suppression resulted in phenotypes that were relevant to the human clinical features of microcephaly and reduced motility in exosc3 morphants. Additionally, in vivo complementation of exosc3 MO phenotypes with either zebrafish or human mRNA harboring the missense mutations found in patients failed to rescue the phenotype, indicating that these were loss-of-function alleles [64]. Even so, transient in vivo complementation assays are not applicable to every gene. Human genes with an open reading frame (ORF) larger than ~6 kb are challenging to transcribe in vitro, likely explaining why large genes such as NBEAL2, encoding neurobeachin-like 2, the novel genetic cause for gray platelet syndrome (ORF of 8.2 kb), was shown to cause a relevant thrombocytopenia phenotype in zebrafish morphants, but the missense mutations identified in patients were not tested [65].
Figure 1. Zebrafish models of human genetic disease.
A. Mutations in pontocerebellar hypoplasia are caused by EXOSC3 and in vivo complementation studies in zebrafish recapitulate the brain phenotypes observed in patients and demonstrate that missense mutations are functional null variants. Left panel, neuroimaging of affected individuals (A1–A4; top row) and control images (A5–A8; bottom row); right panel, whole-mount in situ hybridization in splice blocking MO-injected embryos in lateral view (A9; inset- dorsal view, with rostral to the left) demonstrated diminished expression of dorsal hindbrain progenitor-specific marker atoh1a and cerebellar-specific marker pvalb7 (A10; quantification). Images reproduced from Wan et al 2012. Nat Genet. 44: 704–708.
B. Congenital abnormalities of the kidney and urinary tract are caused by haploinsufficiency of DSTYK. Left, panels B1-B3, with hypoplasia of the left kidney (Panel B1, kidneys outlined by dashed lines), bilateral hydronephrosis (Panel B2, arrows) caused by ureteropelvic junction obstruction detected at birth, and hydronephrosis only of the left kidney (Panel B3, arrow) caused by ureteropelvic junction obstruction. The intravenous pyelogram in Panel B4 shows blunting of fornices on the right side (white arrow); right, MO-induced knockdown of dstyk embryos, live lateral images show absence of the patent pronephric duct opening (arrows). Images reproduced from Sanna-Cherchi et al 2013. NEJM. 369: 621–629.
C. Adult-onset limb-girdle muscular dystrophy is caused by dominant negative mutations in DNAJB6. C1, Transmission electron microscopy showed early disruption of Z-disks (arrows; left) and autophagic pathology (right) in LGMD1D; C2-C10, Lateral views of zebrafish embryos 2 days post fertilization subjected to whole-mount immunofluorescence staining of slow myosin heavy chain display myofiber abnormalities (arrows). Images reproduced from Sarparanta et al 2012. Nat Genet. 44: 450–455.
D. PLS3 overexpression exerts a protective effect on SMN1 deletion to rescue motor neuron defects in spinal muscular atrophy (SMA). D1, pedigrees of SMA-discordant families showing unaffected (gray) and affected (black) SMN1-deleted siblings; D2, Lateral view of zebrafish embryos injected with control MO, smn MO, PLS3 RNA, and smn MO + PLS3 RNA. Motor axons were visualized with znp1 antibody at 36 hours post fertilization and show rescue of smn MO with PLS3 RNA. Images reproduced from Oprea et al 2008. Science. 320: 524–527.
E. SCRIB and PUF60 suppression drive the multisystemic phenotypes of the 8q24.3 copy number variant. E1, photographs five individuals with the 8q23.4 CNV show craniofacial abnormalities and microcephaly; E2, Lateral and dorsal views of control and scrib or puf60a MO-injected embryos at 5 dpf show head size and craniofacial defects observed in affected individuals. Images reproduced from Daubert et al 2013. Am J Hum Genet. 93: 798–811. LOF, loss-of-function; DN, dominant negative.
Dominant disorders
In contrast to recessive disorders, in which the allele effect is typically loss-of-function, traits that segregate under an autosomal dominant inheritance pattern are the result of either a haploinsufficiency, dominant negative, or gain-of-function mechanism. For some dominant pediatric-onset disorders, the genetic evidence of a heterozygous null variant segregating in a large pedigree with fully penetrant disease is sufficient to suggest that the direction of allele effect is haploinsufficiency, and these predictions have been confirmed in zebrafish through MO-induced gene suppression. For example, dilated cardiomyopathy is caused by nonsense, splice-site or missense mutations in the gene encoding heat shock protein co-chaperone BCL2-associated athanogene 3 (BAG3) and gene suppression results in similar cardiac phenotypes in zebrafish embryos [66]. Similarly, congenital abnormalities of the kidney and the urinary tract (CAKUT) associated with a loss-of-function splice site mutation segregating in a dominant pedigree were identified in the dual serine-threonine kinase encoded by DSTYK (Figure 1B); the human phenotypes were recapitulated in dstyk morphant embryos [67]. Transient experiments in zebrafish embryos can also determine allele pathogenicity and capture a dominant negative direction of effect that is isoform-specific. This is exemplified by the transient functional studies of missense mutations in the co-chaperone protein, DNAJB6, associated recently with adult-onset limb girdle muscular dystrophy (Figure 1C) [68]. Co-injection of mutant DNAJB6 mRNA in the presence of equivalent amounts of wild-type transcript resulted in myofiber defects in zebrafish embryos; injection of increasing amounts wild-type mRNA with a fixed concentration of mutant resulted in a phenotypic rescue, indicating the dominant toxicity of the mutant alleles. Given the nature of these mutations (deleterious in heterozygosity), the use of zebrafish for dissecting dominant disorders will likely remain restricted to transient MO- and mRNA-based studies until the development of more sophisticated conditional gene suppression/expression techniques in zebrafish.
De novo variants
Variants that arise de novo as a product of either germline mosaicism or early developmental DNA replication errors are significant contributors to the human mutational burden [69]. Similar to variants that underscore autosomal dominant disorders, de novo changes may give rise to clinical phenotypes produced from falling below a gene dosage threshold, dominant negative effects, or acquisition of a novel function. As such, an unbiased approach towards dissecting the direction of de novo allele effect is critical once physiological relevance has been determined. For instance, transient approaches have been carried out in zebrafish to investigate de novo missense mutations in CACNA1C, encoding the voltage-gated calcium channel Cav1.2, in the pathophysiology of Timothy syndrome (TS), a pediatric disorder characterized by cardiac arrhythmias, syndactyly, and craniofacial abnormalities. Ectopic expression of mutant mRNA and suppression of cacna1c in zebrafish embryos not only revealed that the mutation confers a gain-of-function effect, but also demonstrated a novel role for Cav1.2 in the non-excitable cells of the developing jaw [70].
CNVs
Frequently arising de novo, copy number variants (CNVs) likewise represent a significant molecular basis for human genetic disease [71]. These variations in genomic structure, can range in size from a few thousand to millions of base pairs, are not identifiable by conventional chromosomal banding, and can encompass from one to hundreds of genes [72]. Although genotype-phenotype correlations among affected individuals with overlapping CNVs can assist in narrowing specific genetic drivers, CNVs have been historically intractable to functional interpretation in animal models, with sparse reports of human CNVs being modeled in the mouse [73]. Zebrafish models have emerged recently as powerful tools to dissect both recurrent and non-recurrent CNVs. First, systematic zebrafish modeling of the 29 genes in the recurrent reciprocal 16p11.2 duplication/deletion CNV— associated with a range of neurocognitive defects— found the main driver of the neuroanatomical phenotypes to be KCTD13, causing mirrored macro- and microcephaly upon suppression or overexpression in zebrafish, respectively [74]. Second, MO-induced suppression of three genes in the 8q24.3 non-recurrent deletion CNV in zebrafish embryos revealed that the planar cell polarity effector SCRIB, and the splicing factor PUF60 could be linked to distinct aspects of the renal, short stature, coloboma, and cardiac phenotypes observed in five individuals with overlapping microdeletions at this locus (Figure 1E) [75].
Second-site modifiers
The demonstration of second-site phenotype modification in primarily recessive human genetic disease has been fuelled by the use of in vivo assays in zebrafish. The ciliopathies, disorders underscored by dysfunction of the primary cilium, have been causally linked with over 50 different loci, can give rise to a constellation of human phenotypes, and have been an ideal system to study epistasis [76]. The recent dissection of the genetic architecture of Bardet-Biedl syndrome (BBS), a ciliopathy hallmarked by retinal degeneration, obesity, postaxial polydactyly, renal abnormalities, and intellectual disability, 1) informed the pathogenic potential of missense BBS alleles contributing to the disorder (null, hypomorphic, or dominant negative); 2) revealed the surprising contribution of dominant negative alleles in oligogenic pedigrees with BBS; and 3) provided sensitivity (98%) and specificity (82%) metrics for the zebrafish in vivo complementation assay to predict allele pathogenicity [77]. Transient zebrafish in vivo complementation assays have similarly been used to identify RPGRIP1L A229T as a modulator of retinal endophenotypes [78], RET as a modifier of Hirschsprung phenotypes in BBS [79], and TTC21B as a frequent contributor to mutational load in ciliopathies [80]. Second-site modification phenomena are not unique to the ciliopathies; for example, overexpression of plastin 3 (PLS3) to mimic the gene expression in unaffected individuals, improved the axon length and growth defects associated with SMN1 deletion in spinal muscular atrophy (SMA). The interaction of these two genes was shown, in part, through modeling of SMA genotype and phenotype correlates in zebrafish embryos (Figure 1D) [81].
Complex traits
Genome wide association studies (GWAS) alone have been hampered by an inability to connect risk association to genes and underlying mechanism. However, the zebrafish has emerged as a tool to dissect genes at or near loci that confer significant risk for the complex trait under investigation. Rare alleles in GWAS hits have been more straightforward to dissect, since the strategy has been similar to that of Mendelian traits. For example, a combination of in vitro analysis of enzyme stability and secretion and vascular integrity in the retina in zebrafish embryos demonstrated a functional role for a rare allele in the gene encoding complement factor I (CFI), thus providing direct evidence for a loss of function role of CFI in AMD [82]. Zebrafish have also been used in the absence of candidate coding changes in GWAS-identified loci. In one example, Liu et al assessed the physiological relevance of candidate genes identified from a GWAS of chronic kidney disease among African Americans populations; these efforts identified KCNQ1 as a functionally relevant candidate due to the glomerular filtration defects observed in kcnq1 morphants [83]. Similarly, functional validation of loci associated with platelet count in cohorts of European ancestry yielded eleven novel genes implicated Danio rerio blood cell formation [84]. Although numerous questions remain regarding the combinatorial effects of GWAS hits, their mechanistic basis for conferring risk, and the physiological relevance of significantly associated sites, both coding and non-coding, the zebrafish offers a tractable tool to begin to dissect existing GWAS data.
Adult Onset Disease
The majority of zebrafish models discussed so far have been used to understand the role of genes and alleles in pediatric and congenital disorders, in large part because transient MO and mRNA analysis is possible only during development. Nonetheless, given the correct tools and appropriate assays, this model organism is also useful for the study of adult onset disorders. AMD was one example described above, the utility of the model being extracted from the ability to model vascular integrity in zebrafish embryos, a phenotype relevant to AMD pathology. There are numerous other examples as well. In particular, phenotypic proxies of human neurodegenerative diseases such as schizophrenia, Huntington, Parkinson, and Alzheimer disease, have been used to examine the role various genes in these diseases (reviewed extensively in [85] and [86]), although in almost all cases these studies involved genetic mutants and/or stable transgenes. In one example, stable transgenic zebrafish expressing human 4-repeat Tau showed Tau accumulation within neuronal cell bodies axons in neurons throughout the adult brain, resembling neurofibrillary tangles [87]. Some defects, such as behavioral phenotypes, are not immediately observable through anatomic or histological methods, making it necessary to employ more sensitive methods of analysis. To this extent, assays for memory and learning impairment [88] and conditioned avoidance [89] allow for quantifiable testing of subtle phenotypes in adult fish.
Conclusion: The road ahead
In the study of human genetics, animal models have provided insight into the genetics and pathophysiology. The efficacy of a model organism always hinges on whether that organism models appropriately the target pathology of humans and whether the experiments necessary to provide burden of proof are tractable and not cost-prohibitive. In the context of human and medical genomics, we anticipate that the entire spectrum of model organisms will continue to be used. Nonetheless, it is clear that models such as zebrafish and possibly other similar organisms, such as Medaka or Xenopus that offer transparency, low cost and the ability to manipulate their genome efficiently will gain a prominent role as the community strives to model thousands of candidate disease-associated genes and alleles.
A key requirement towards the widespread use of the zebrafish to determine pathogenicity of alleles identified by WES/WGS moving forward will be to a) improve throughput of physiologically-relevant models of human disease (through the rapid generation of in vivo complementation assays); and b) automated phenotyping and image analysis. While MOs may still be a tractable option for the evaluation of early developmental phenotypes, the recent ability to generate knock-in mutants [90], conditional zebrafish mutants [91], and multiplexed mutants [92] holds great promise towards accelerating disease modeling throughput. Equally as important, some platforms for high throughput phenotypic screening of zebrafish larvae have been developed recently [93]. For example, one system captures hundreds of three-dimensional morphological features with speed and accuracy, clustering quantitative phenotypic signatures so that multiple phenotypes can be detected and classified simultaneously. Nonetheless, this system is limited to bright field images, rendering marker analysis (by RNA in situ or antibody staining) difficult. Given that embryo phenotyping represents the most significant bottleneck in scaling the use of zebrafish in human genomics, there is an acute need to develop additional transgenic reporter lines to assist with visualization of cellular and anatomical structures of interest [94], as well as imaging and embryo manipulation technologies further.
It is also important to note that zebrafish analysis, like all other genetic and molecular biology tools, has its limitations. MO studies can sometimes generate conflicting data, especially in the context of early developmental phenotypes that are most sensitive to toxic effects. Rescue studies and a minimum of two MOs per gene tested, when possible, are essential to validate findings. Similarly, multiple independent lines from genome editing experiments will need to be studied to ensure that the phenotypes observed are driven by the engineered mutation, not an off-site introduced allele. Moreover, not all human genes and alleles are modelable; some 25–30% of the human transcriptome is not present in zebrafish, while other genes can be difficult to model because of divergent functions or extreme dosage sensitivity (especially transcription factors) rendering them experimentally difficult. Further, for some disorders (e.g. pulmonary fibrosis) there cannot be a credible phenotypic surrogate and distant surrogates might lead to incorrect conclusions. Finally, most zebrafish studies to date have focused on coding variation; modeling non-coding variation is significantly more taxing, yet remain important, not least because regulatory regions are likely to be enriched for alleles that drive GWAS signals [95]. This work is possible, once we recognize that evolutionary constraints might render some data uninterpretable. For example, testing multiple sequences located within a 50 kb block of the regulatory domain of IRX3 (certain variants of which are associated with obesity in humans) resulted in transgenic zebrafish with expression in pancreas [96]; knockdown of irx3 in zebrafish reduced the number of pancreatic beta-cells. However, given current designs, it will be difficult to execute such experiments at the throughout required to address the needs of the human genetics community.
Despite these limitations, modeling human variation in zebrafish embryos has been a significant contributor towards dissecting the causality of genes and alleles in human genetic disorders. Moreover, the development of human disease models will serve as a platform for the discovery of novel therapeutic paradigms by employing high-throughput small molecule screening approaches (reviewed extensively elsewhere [97, 98]). Moving forward, we anticipate that the community, through the combinatorial use of all the tools discussed here, will saturate in the coming years a significant fraction of the morbid human genome. Ultimately, we imagine that such studies will inform the design of improved computational algorithms, probably through the training of thousands of human alleles tested in vivo, which will in turn represent the next inflection point in human and medical genomics.
Highlights.
Interpretation of rare and novel variants is a current challenge in human genetics
Zebrafish are a model system to study the pathogenicity of human alleles
Conserved evolutionary and physiological features make zebrafish an ideal model
Zebrafish are suitable for moderate to high throughput assays of human disease
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- 1.Lejeune J, Gautier M, Turpin R. [Study of somatic chromosomes from 9 mongoloid children] Comptes rendus hebdomadaires des seances de l'Academie des sciences. 1959;248:1721–1722. [PubMed] [Google Scholar]
- 2.Collins FS. Positional cloning: let's not call it reverse anymore. Nature genetics. 1992;1:3–6. doi: 10.1038/ng0492-3. [DOI] [PubMed] [Google Scholar]
- 3.Golzio C, Katsanis N. Genetic architecture of reciprocal CNVs. Current opinion in genetics & development. 2013;23:240–248. doi: 10.1016/j.gde.2013.04.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Katsanis SH, Katsanis N. Molecular genetic testing and the future of clinical genomics. Nature reviews. Genetics. 2013;14:415–426. doi: 10.1038/nrg3493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Genomes Project C, Abecasis GR, Altshuler D, Auton A, Brooks LD, Durbin RM, Gibbs RA, Hurles ME, McVean GA. A map of human genome variation from population-scale sequencing. Nature. 2010;467:1061–1073. doi: 10.1038/nature09534. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Tennessen JA, Bigham AW, O'Connor TD, Fu W, Kenny EE, Gravel S, McGee S, Do R, Liu X, Jun G, Kang HM, Jordan D, Leal SM, Gabriel S, Rieder MJ, Abecasis G, Altshuler D, Nickerson DA, Boerwinkle E, Sunyaev S, Bustamante CD, Bamshad MJ, Akey JM, Broad GO, Seattle GO, Project NES. Evolution and functional impact of rare coding variation from deep sequencing of human exomes. Science. 2012;337:64–69. doi: 10.1126/science.1219240. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Castellana S, Mazza T. Congruency in the prediction of pathogenic missense mutations: state-of-the-art web-based tools. Briefings in bioinformatics. 2013;14:448–459. doi: 10.1093/bib/bbt013. [DOI] [PubMed] [Google Scholar]
- 8.Aitman TJ, Boone C, Churchill GA, Hengartner MO, Mackay TF, Stemple DL. The future of model organisms in human disease research. Nature reviews. Genetics. 2011;12:575–582. doi: 10.1038/nrg3047. [DOI] [PubMed] [Google Scholar]
- 9.Niederriter AR, Davis EE, Golzio C, Oh EC, Tsai IC, Katsanis N. In vivo modeling of the morbid human genome using Danio rerio. Journal of visualized experiments : JoVE. 2013:e50338. doi: 10.3791/50338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Mouse Genome Sequencing C, Waterston RH, Lindblad-Toh K, Birney E, Rogers J, Abril JF, Agarwal P, Agarwala R, Ainscough R, Alexandersson M, An P, Antonarakis SE, Attwood J, Baertsch R, Bailey J, Barlow K, Beck S, Berry E, Birren B, Bloom T, Bork P, Botcherby M, Bray N, Brent MR, Brown DG, Brown SD, Bult C, Burton J, Butler J, Campbell RD, Carninci P, Cawley S, Chiaromonte F, Chinwalla AT, Church DM, Clamp M, Clee C, Collins FS, Cook LL, Copley RR, Coulson A, Couronne O, Cuff J, Curwen V, Cutts T, Daly M, David R, Davies J, Delehaunty KD, Deri J, Dermitzakis ET, Dewey C, Dickens NJ, Diekhans M, Dodge S, Dubchak I, Dunn DM, Eddy SR, Elnitski L, Emes RD, Eswara P, Eyras E, Felsenfeld A, Fewell GA, Flicek P, Foley K, Frankel WN, Fulton LA, Fulton RS, Furey TS, Gage D, Gibbs RA, Glusman G, Gnerre S, Goldman N, Goodstadt L, Grafham D, Graves TA, Green ED, Gregory S, Guigo R, Guyer M, Hardison RC, Haussler D, Hayashizaki Y, Hillier LW, Hinrichs A, Hlavina W, Holzer T, Hsu F, Hua A, Hubbard T, Hunt A, Jackson I, Jaffe DB, Johnson LS, Jones M, Jones TA, Joy A, Kamal M, Karlsson EK, Karolchik D, Kasprzyk A, Kawai J, Keibler E, Kells C, Kent WJ, Kirby A, Kolbe DL, Korf I, Kucherlapati RS, Kulbokas EJ, Kulp D, Landers T, Leger JP, Leonard S, Letunic I, Levine R, Li J, Li M, Lloyd C, Lucas S, Ma B, Maglott DR, Mardis ER, Matthews L, Mauceli E, Mayer JH, McCarthy M, McCombie WR, McLaren S, McLay K, McPherson JD, Meldrim J, Meredith B, Mesirov JP, Miller W, Miner TL, Mongin E, Montgomery KT, Morgan M, Mott R, Mullikin JC, Muzny DM, Nash WE, Nelson JO, Nhan MN, Nicol R, Ning Z, Nusbaum C, O'Connor MJ, Okazaki Y, Oliver K, Overton-Larty E, Pachter L, Parra G, Pepin KH, Peterson J, Pevzner P, Plumb R, Pohl CS, Poliakov A, Ponce TC, Ponting CP, Potter S, Quail M, Reymond A, Roe BA, Roskin KM, Rubin EM, Rust AG, Santos R, Sapojnikov V, Schultz B, Schultz J, Schwartz MS, Schwartz S, Scott C, Seaman S, Searle S, Sharpe T, Sheridan A, Shownkeen R, Sims S, Singer JB, Slater G, Smit A, Smith DR, Spencer B, Stabenau A, Stange-Thomann N, Sugnet C, Suyama M, Tesler G, Thompson J, Torrents D, Trevaskis E, Tromp J, Ucla C, Ureta-Vidal A, Vinson JP, Von Niederhausern AC, Wade CM, Wall M, Weber RJ, Weiss RB, Wendl MC, West AP, Wetterstrand K, Wheeler R, Whelan S, Wierzbowski J, Willey D, Williams S, Wilson RK, Winter E, Worley KC, Wyman D, Yang S, Yang SP, Zdobnov EM, Zody MC, Lander ES. Initial sequencing and comparative analysis of the mouse genome. Nature. 2002;420:520–562. doi: 10.1038/nature01262. [DOI] [PubMed] [Google Scholar]
- 11.Capecchi MR. Gene targeting in mice: functional analysis of the mammalian genome for the twenty-first century. Nature reviews. Genetics. 2005;6:507–512. doi: 10.1038/nrg1619. [DOI] [PubMed] [Google Scholar]
- 12.Devoy A, Bunton-Stasyshyn RK, Tybulewicz VL, Smith AJ, Fisher EM. Genomically humanized mice: technologies and promises. Nature reviews. Genetics. 2012;13:14–20. doi: 10.1038/nrg3116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Antoshechkin I, Sternberg PW. The versatile worm: genetic and genomic resources for Caenorhabditis elegans research. Nature reviews. Genetics. 2007;8:518–532. doi: 10.1038/nrg2105. [DOI] [PubMed] [Google Scholar]
- 14.Wijshake T, Baker DJ, van de Sluis B. Endonucleases: new tools to edit the mouse genome. Biochim. Biophys. Acta. 2014 doi: 10.1016/j.bbadis.2014.04.020. [DOI] [PubMed] [Google Scholar]
- 15.Lander ES, Linton LM, Birren B, Nusbaum C, Zody MC, Baldwin J, Devon K, Dewar K, Doyle M, FitzHugh W, Funke R, Gage D, Harris K, Heaford A, Howland J, Kann L, Lehoczky J, LeVine R, McEwan P, McKernan K, Meldrim J, Mesirov JP, Miranda C, Morris W, Naylor J, Raymond C, Rosetti M, Santos R, Sheridan A, Sougnez C, Stange-Thomann N, Stojanovic N, Subramanian A, Wyman D, Rogers J, Sulston J, Ainscough R, Beck S, Bentley D, Burton J, Clee C, Carter N, Coulson A, Deadman R, Deloukas P, Dunham A, Dunham I, Durbin R, French L, Grafham D, Gregory S, Hubbard T, Humphray S, Hunt A, Jones M, Lloyd C, McMurray A, Matthews L, Mercer S, Milne S, Mullikin JC, Mungall A, Plumb R, Ross M, Shownkeen R, Sims S, Waterston RH, Wilson RK, Hillier LW, McPherson JD, Marra MA, Mardis ER, Fulton LA, Chinwalla AT, Pepin KH, Gish WR, Chissoe SL, Wendl MC, Delehaunty KD, Miner TL, Delehaunty A, Kramer JB, Cook LL, Fulton RS, Johnson DL, Minx PJ, Clifton SW, Hawkins T, Branscomb E, Predki P, Richardson P, Wenning S, Slezak T, Doggett N, Cheng JF, Olsen A, Lucas S, Elkin C, Uberbacher E, Frazier M, Gibbs RA, Muzny DM, Scherer SE, Bouck JB, Sodergren EJ, Worley KC, Rives CM, Gorrell JH, Metzker ML, Naylor SL, Kucherlapati RS, Nelson DL, Weinstock GM, Sakaki Y, Fujiyama A, Hattori M, Yada T, Toyoda A, Itoh T, Kawagoe C, Watanabe H, Totoki Y, Taylor T, Weissenbach J, Heilig R, Saurin W, Artiguenave F, Brottier P, Bruls T, Pelletier E, Robert C, Wincker P, Smith DR, Doucette-Stamm L, Rubenfield M, Weinstock K, Lee HM, Dubois J, Rosenthal A, Platzer M, Nyakatura G, Taudien S, Rump A, Yang H, Yu J, Wang J, Huang G, Gu J, Hood L, Rowen L, Madan A, Qin S, Davis RW, Federspiel NA, Abola AP, Proctor MJ, Myers RM, Schmutz J, Dickson M, Grimwood J, Cox DR, Olson MV, Kaul R, Raymond C, Shimizu N, Kawasaki K, Minoshima S, Evans GA, Athanasiou M, Schultz R, Roe BA, Chen F, Pan H, Ramser J, Lehrach H, Reinhardt R, McCombie WR, de la Bastide M, Dedhia N, Blocker H, Hornischer K, Nordsiek G, Agarwala R, Aravind L, Bailey JA, Bateman A, Batzoglou S, Birney E, Bork P, Brown DG, Burge CB, Cerutti L, Chen HC, Church D, Clamp M, Copley RR, Doerks T, Eddy SR, Eichler EE, Furey TS, Galagan J, Gilbert JG, Harmon C, Hayashizaki Y, Haussler D, Hermjakob H, Hokamp K, Jang W, Johnson LS, Jones TA, Kasif S, Kaspryzk A, Kennedy S, Kent WJ, Kitts P, Koonin EV, Korf I, Kulp D, Lancet D, Lowe TM, McLysaght A, Mikkelsen T, Moran JV, Mulder N, Pollara VJ, Ponting CP, Schuler G, Schultz J, Slater G, Smit AF, Stupka E, Szustakowski J, Thierry-Mieg D, Thierry-Mieg J, Wagner L, Wallis J, Wheeler R, Williams A, Wolf YI, Wolfe KH, Yang SP, Yeh RF, Collins F, Guyer MS, Peterson J, Felsenfeld A, Wetterstrand KA, Patrinos A, Morgan MJ, de Jong P, Catanese JJ, Osoegawa K, Shizuya H, Choi S, Chen, C YJ. International Human Genome Sequencing, Initial sequencing and analysis of the human genome. Nature. 2001;409:860–921. doi: 10.1038/35057062. [DOI] [PubMed] [Google Scholar]
- 16.Streisinger G, Walker C, Dower N, Knauber D, Singer F. Production of clones of homozygous diploid zebra fish (Brachydanio rerio) Nature. 1981;291:293–296. doi: 10.1038/291293a0. [DOI] [PubMed] [Google Scholar]
- 17.Amores A, Force A, Yan YL, Joly L, Amemiya C, Fritz A, Ho RK, Langeland J, Prince V, Wang YL, Westerfield M, Ekker M, Postlethwait JH. Zebrafish hox clusters and vertebrate genome evolution. Science. 1998;282:1711–1714. doi: 10.1126/science.282.5394.1711. [DOI] [PubMed] [Google Scholar]
- 18.Meyer A, Schartl M. Gene and genome duplications in vertebrates: the one-to-four (-to-eight in fish) rule and the evolution of novel gene functions. Current opinion in cell biology. 1999;11:699–704. doi: 10.1016/s0955-0674(99)00039-3. [DOI] [PubMed] [Google Scholar]
- 19.Postlethwait JH, Yan YL, Gates MA, Horne S, Amores A, Brownlie A, Donovan A, Egan ES, Force A, Gong Z, Goutel C, Fritz A, Kelsh R, Knapik E, Liao E, Paw B, Ransom D, Singer A, Thomson M, Abduljabbar TS, Yelick P, Beier D, Joly JS, Larhammar D, Rosa F, Westerfield M, Zon LI, Johnson SL, Talbot WS. Vertebrate genome evolution and the zebrafish gene map. Nature genetics. 1998;18:345–349. doi: 10.1038/ng0498-345. [DOI] [PubMed] [Google Scholar]
- 20.Howe K, Clark MD, Torroja CF, Torrance J, Berthelot C, Muffato M, Collins JE, Humphray S, McLaren K, Matthews L, McLaren S, Sealy I, Caccamo M, Churcher C, Scott C, Barrett JC, Koch R, Rauch GJ, White S, Chow W, Kilian B, Quintais LT, Guerra-Assuncao JA, Zhou Y, Gu Y, Yen J, Vogel JH, Eyre T, Redmond S, Banerjee R, Chi J, Fu B, Langley E, Maguire SF, Laird GK, Lloyd D, Kenyon E, Donaldson S, Sehra H, Almeida-King J, Loveland J, Trevanion S, Jones M, Quail M, Willey D, Hunt A, Burton J, Sims S, McLay K, Plumb B, Davis J, Clee C, Oliver K, Clark R, Riddle C, Elliot D, Threadgold G, Harden G, Ware D, Mortimore B, Kerry G, Heath P, Phillimore B, Tracey A, Corby N, Dunn M, Johnson C, Wood J, Clark S, Pelan S, Griffiths G, Smith M, Glithero R, Howden P, Barker N, Stevens C, Harley J, Holt K, Panagiotidis G, Lovell J, Beasley H, Henderson C, Gordon D, Auger K, Wright D, Collins J, Raisen C, Dyer L, Leung K, Robertson L, Ambridge K, Leongamornlert D, McGuire S, Gilderthorp R, Griffiths C, Manthravadi D, Nichol S, Barker G, Whitehead S, Kay M, Brown J, Murnane C, Gray E, Humphries M, Sycamore N, Barker D, Saunders D, Wallis J, Babbage A, Hammond S, Mashreghi-Mohammadi M, Barr L, Martin S, Wray P, Ellington A, Matthews N, Ellwood M, Woodmansey R, Clark G, Cooper J, Tromans A, Grafham D, Skuce C, Pandian R, Andrews R, Harrison E, Kimberley A, Garnett J, Fosker N, Hall R, Garner P, Kelly D, Bird C, Palmer S, Gehring I, Berger A, Dooley CM, Ersan-Urun Z, Eser C, Geiger H, Geisler M, Karotki L, Kirn A, Konantz J, Konantz M, Oberlander M, Rudolph-Geiger S, Teucke M, Osoegawa K, Zhu B, Rapp A, Widaa S, Langford C, Yang F, Carter NP, Harrow J, Ning Z, Herrero J, Searle SM, Enright A, Geisler R, Plasterk RH, Lee C, Westerfield M, de Jong PJ, Zon LI, Postlethwait JH, Nusslein-Volhard C, Hubbard TJ, Roest Crollius H, Rogers J, Stemple DL, Begum S, Lloyd C, Lanz C, Raddatz G, Schuster SC. The zebrafish reference genome sequence and its relationship to the human genome. Nature. 2013;496:498–503. doi: 10.1038/nature12111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Howe DG, Bradford YM, Conlin T, Eagle AE, Fashena D, Frazer K, Knight J, Mani P, Martin R, Moxon SA, Paddock H, Pich C, Ramachandran S, Ruef BJ, Ruzicka L, Schaper K, Shao X, Singer A, Sprunger B, Van Slyke CE, Westerfield, ZFIN M. the Zebrafish Model Organism Database: increased support for mutants and transgenics. Nucleic acids research. 2013;41:D854–D860. doi: 10.1093/nar/gks938. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Lawson ND, Wolfe SA. Forward and reverse genetic approaches for the analysis of vertebrate development in the zebrafish. Developmental cell. 2011;21:48–64. doi: 10.1016/j.devcel.2011.06.007. [DOI] [PubMed] [Google Scholar]
- 23.Chakrabarti S, Streisinger G, Singer F, Walker C. Frequency of gamma-Ray Induced Specific Locus and Recessive Lethal Mutations in Mature Germ Cells of the Zebrafish, BRACHYDANIO RERIO. Genetics. 1983;103:109–123. doi: 10.1093/genetics/103.1.109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Streisinger G. Extrapolations from species to species and from various cell types in assessing risks from chemical mutagens. Mutation research. 1983;114:93–105. doi: 10.1016/0165-1110(83)90021-0. [DOI] [PubMed] [Google Scholar]
- 25.Mullins MC, Hammerschmidt M, Haffter P, Nusslein-Volhard C. Large-scale mutagenesis in the zebrafish: in search of genes controlling development in a vertebrate. Current biology : CB. 1994;4:189–202. doi: 10.1016/s0960-9822(00)00048-8. [DOI] [PubMed] [Google Scholar]
- 26.Solnica-Krezel L, Schier AF, Driever W. Efficient recovery of ENU-induced mutations from the zebrafish germline. Genetics. 1994;136:1401–1420. doi: 10.1093/genetics/136.4.1401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Hammerschmidt M, Pelegri F, Mullins MC, Kane DA, Brand M, van Eeden FJ, Furutani-Seiki M, Granato M, Haffter P, Heisenberg CP, Jiang YJ, Kelsh RN, Odenthal J, Warga RM, Nusslein-Volhard C. Mutations affecting morphogenesis during gastrulation and tail formation in the zebrafish, Danio rerio. Development. 1996;123:143–151. doi: 10.1242/dev.123.1.143. [DOI] [PubMed] [Google Scholar]
- 28.Solnica-Krezel L, Stemple DL, Mountcastle-Shah E, Rangini Z, Neuhauss SC, Malicki J, Schier AF, Stainier DY, Zwartkruis F, Abdelilah S, Driever W. Mutations affecting cell fates and cellular rearrangements during gastrulation in zebrafish. Development. 1996;123:67–80. doi: 10.1242/dev.123.1.67. [DOI] [PubMed] [Google Scholar]
- 29.van Eeden FJ, Granato M, Schach U, Brand M, Furutani-Seiki M, Haffter P, Hammerschmidt M, Heisenberg CP, Jiang YJ, Kane DA, Kelsh RN, Mullins MC, Odenthal J, Warga RM, Allende ML, Weinberg ES, Nusslein-Volhard C. Mutations affecting somite formation and patterning in the zebrafish, Danio rerio. Development. 1996;123:153–164. doi: 10.1242/dev.123.1.153. [DOI] [PubMed] [Google Scholar]
- 30.Brand M, Heisenberg CP, Jiang YJ, Beuchle D, Lun K, Furutani-Seiki M, Granato M, Haffter P, Hammerschmidt M, Kane DA, Kelsh RN, Mullins MC, Odenthal J, van Eeden FJ, Nusslein-Volhard C. Mutations in zebrafish genes affecting the formation of the boundary between midbrain and hindbrain. Development. 1996;123:179–190. doi: 10.1242/dev.123.1.179. [DOI] [PubMed] [Google Scholar]
- 31.Jiang YJ, Brand M, Heisenberg CP, Beuchle D, Furutani-Seiki M, Kelsh RN, Warga RM, Granato M, Haffter P, Hammerschmidt M, Kane DA, Mullins MC, Odenthal J, van Eeden FJ, Nusslein-Volhard C. Mutations affecting neurogenesis and brain morphology in the zebrafish, Danio rerio. Development. 1996;123:205–216. doi: 10.1242/dev.123.1.205. [DOI] [PubMed] [Google Scholar]
- 32.Schier AF, Neuhauss SC, Harvey M, Malicki J, Solnica-Krezel L, Stainier DY, Zwartkruis F, Abdelilah S, Stemple DL, Rangini Z, Yang H, Driever W. Mutations affecting the development of the embryonic zebrafish brain. Development. 1996;123:165–178. doi: 10.1242/dev.123.1.165. [DOI] [PubMed] [Google Scholar]
- 33.Stainier DY, Fouquet B, Chen JN, Warren KS, Weinstein BM, Meiler SE, Mohideen MA, Neuhauss SC, Solnica-Krezel L, Schier AF, Zwartkruis F, Stemple DL, Malicki J, Driever W, Fishman MC. Mutations affecting the formation and function of the cardiovascular system in the zebrafish embryo. Development. 1996;123:285–292. doi: 10.1242/dev.123.1.285. [DOI] [PubMed] [Google Scholar]
- 34.Neuhauss SC, Solnica-Krezel L, Schier AF, Zwartkruis F, Stemple DL, Malicki J, Abdelilah S, Stainier DY, Driever W. Mutations affecting craniofacial development in zebrafish. Development. 1996;123:357–367. doi: 10.1242/dev.123.1.357. [DOI] [PubMed] [Google Scholar]
- 35.Piotrowski T, Schilling TF, Brand M, Jiang YJ, Heisenberg CP, Beuchle D, Grandel H, van Eeden FJ, Furutani-Seiki M, Granato M, Haffter P, Hammerschmidt M, Kane DA, Kelsh RN, Mullins MC, Odenthal J, Warga RM, Nusslein-Volhard C. Jaw and branchial arch mutants in zebrafish II: anterior arches and cartilage differentiation. Development. 1996;123:345–356. doi: 10.1242/dev.123.1.345. [DOI] [PubMed] [Google Scholar]
- 36.Schilling TF, Piotrowski T, Grandel H, Brand M, Heisenberg CP, Jiang YJ, Beuchle D, Hammerschmidt M, Kane DA, Mullins MC, van Eeden FJ, Kelsh RN, Furutani-Seiki M, Granato M, Haffter P, Odenthal J, Warga RM, Trowe T, Nusslein-Volhard C. Jaw and branchial arch mutants in zebrafish I: branchial arches. Development. 1996;123:329–344. doi: 10.1242/dev.123.1.329. [DOI] [PubMed] [Google Scholar]
- 37.Lang MR, Lapierre LA, Frotscher M, Goldenring JR, Knapik EW. Secretory COPII coat component Sec23a is essential for craniofacial chondrocyte maturation. Nature genetics. 2006;38:1198–1203. doi: 10.1038/ng1880. [DOI] [PubMed] [Google Scholar]
- 38.Boyadjiev SA, Fromme JC, Ben J, Chong SS, Nauta C, Hur DJ, Zhang G, Hamamoto S, Schekman R, Ravazzola M, Orci L, Eyaid W. Cranio-lenticulo-sutural dysplasia is caused by a SEC23A mutation leading to abnormal endoplasmic-reticulum-to-Golgi trafficking. Nature genetics. 2006;38:1192–1197. doi: 10.1038/ng1876. [DOI] [PubMed] [Google Scholar]
- 39.Obholzer N, Swinburne IA, Schwab E, Nechiporuk AV, Nicolson T, Megason SG. Rapid positional cloning of zebrafish mutations by linkage and homozygosity mapping using whole-genome sequencing. Development. 2012;139:4280–4290. doi: 10.1242/dev.083931. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Ryan S, Willer J, Marjoram L, Bagwell J, Mankiewicz J, Leshchiner I, Goessling W, Bagnat M, Katsanis N. Rapid identification of kidney cyst mutations by whole exome sequencing in zebrafish. Development. 2013;140:4445–4451. doi: 10.1242/dev.101170. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Leshchiner I, Alexa K, Kelsey P, Adzhubei I, Austin-Tse CA, Cooney JD, Anderson H, King MJ, Stottmann RW, Garnaas MK, Ha S, Drummond IA, Paw BH, North TE, Beier DR, Goessling W, Sunyaev SR. Mutation mapping and identification by whole-genome sequencing. Genome research. 2012;22:1541–1548. doi: 10.1101/gr.135541.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Summerton J, Weller D. Morpholino antisense oligomers: design, preparation, and properties. Antisense & nucleic acid drug development. 1997;7:187–195. doi: 10.1089/oli.1.1997.7.187. [DOI] [PubMed] [Google Scholar]
- 43.Nasevicius A, Ekker SC. Effective targeted gene 'knockdown' in zebrafish. Nature genetics. 2000;26:216–220. doi: 10.1038/79951. [DOI] [PubMed] [Google Scholar]
- 44.Niederriter AR, Davis EE, Golzio C, Oh EC, Tsai IC, Katsanis N. In Vivo Modeling of the Morbid Human Genome using Danio rerio. 2013:e50338. doi: 10.3791/50338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Shestopalov IA, Sinha S, Chen JK. Light-controlled gene silencing in zebrafish embryos. Nature chemical biology. 2007;3:650–651. doi: 10.1038/nchembio.2007.30. [DOI] [PubMed] [Google Scholar]
- 46.Eisen JS, Smith JC. Controlling morpholino experiments: don't stop making antisense. Development. 2008;135:1735–1743. doi: 10.1242/dev.001115. [DOI] [PubMed] [Google Scholar]
- 47.Wienholds E, Schulte-Merker S, Walderich B, Plasterk RH. Target-selected inactivation of the zebrafish rag1 gene. Science. 2002;297:99–102. doi: 10.1126/science.1071762. [DOI] [PubMed] [Google Scholar]
- 48.Kettleborough RN, Busch-Nentwich EM, Harvey SA, Dooley CM, de Bruijn E, van Eeden F, Sealy I, White RJ, Herd C, Nijman IJ, Fenyes F, Mehroke S, Scahill C, Gibbons R, Wali N, Carruthers S, Hall A, Yen J, Cuppen E, Stemple DL. A systematic genome-wide analysis of zebrafish protein-coding gene function. Nature. 2013;496:494–497. doi: 10.1038/nature11992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Wang D, Jao LE, Zheng N, Dolan K, Ivey J, Zonies S, Wu X, Wu K, Yang H, Meng Q, Zhu Z, Zhang B, Lin S, Burgess SM. Efficient genome-wide mutagenesis of zebrafish genes by retroviral insertions. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:12428–12433. doi: 10.1073/pnas.0705502104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Sivasubbu S, Balciunas D, Amsterdam A, Ekker SC. Insertional mutagenesis strategies in zebrafish. Genome biology. 2007;8(Suppl 1):S9. doi: 10.1186/gb-2007-8-s1-s9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Urnov FD, Rebar EJ, Holmes MC, Zhang HS, Gregory PD. Genome editing with engineered zinc finger nucleases. Nature reviews. Genetics. 2010;11:636–646. doi: 10.1038/nrg2842. [DOI] [PubMed] [Google Scholar]
- 52.Doyon Y, McCammon JM, Miller JC, Faraji F, Ngo C, Katibah GE, Amora R, Hocking TD, Zhang L, Rebar EJ, Gregory PD, Urnov FD, Amacher SL. Heritable targeted gene disruption in zebrafish using designed zinc-finger nucleases. Nature biotechnology. 2008;26:702–708. doi: 10.1038/nbt1409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Meng X, Noyes MB, Zhu LJ, Lawson ND, Wolfe SA. Targeted gene inactivation in zebrafish using engineered zinc-finger nucleases. Nature biotechnology. 2008;26:695–701. doi: 10.1038/nbt1398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Bedell VM, Wang Y, Campbell JM, Poshusta TL, Starker CG, Krug RG, 2nd, Tan W, Penheiter SG, Ma AC, Leung AY, Fahrenkrug SC, Carlson DF, Voytas DF, Clark KJ, Essner JJ, Ekker SC. In vivo genome editing using a high-efficiency TALEN system. Nature. 2012;491:114–118. doi: 10.1038/nature11537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Hwang WY, Fu Y, Reyon D, Maeder ML, Tsai SQ, Sander JD, Peterson RT, Yeh JR, Joung JK. Efficient genome editing in zebrafish using a CRISPR-Cas system. Nature biotechnology. 2013;31:227–229. doi: 10.1038/nbt.2501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Chang N, Sun C, Gao L, Zhu D, Xu X, Zhu X, Xiong JW, Xi JJ. Genome editing with RNA-guided Cas9 nuclease in zebrafish embryos. Cell research. 2013;23:465–472. doi: 10.1038/cr.2013.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Reiter JF, Alexander J, Rodaway A, Yelon D, Patient R, Holder N, Stainier DY. Gata5 is required for the development of the heart and endoderm in zebrafish. Genes & development. 1999;13:2983–2995. doi: 10.1101/gad.13.22.2983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Padang R, Bagnall RD, Richmond DR, Bannon PG, Semsarian C. Rare non-synonymous variations in the transcriptional activation domains of GATA5 in bicuspid aortic valve disease. Journal of molecular and cellular cardiology. 2012;53:277–281. doi: 10.1016/j.yjmcc.2012.05.009. [DOI] [PubMed] [Google Scholar]
- 59.Wei D, Bao H, Liu XY, Zhou N, Wang Q, Li RG, Xu YJ, Yang YQ. GATA5 loss-of-function mutations underlie tetralogy of fallot. International journal of medical sciences. 2013;10:34–42. doi: 10.7150/ijms.5270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Wei D, Bao H, Zhou N, Zheng GF, Liu XY, Yang YQ. GATA5 loss-of-function mutation responsible for the congenital ventriculoseptal defect. Pediatric cardiology. 2013;34:504–511. doi: 10.1007/s00246-012-0482-6. [DOI] [PubMed] [Google Scholar]
- 61.Hjeij R, Lindstrand A, Francis R, Zariwala MA, Liu X, Li Y, Damerla R, Dougherty GW, Abouhamed M, Olbrich H, Loges NT, Pennekamp P, Davis EE, Carvalho CM, Pehlivan D, Werner C, Raidt J, Kohler G, Haffner K, Reyes-Mugica M, Lupski JR, Leigh MW, Rosenfeld M, Morgan LC, Knowles MR, Lo CW, Katsanis N, Omran H. ARMC4 mutations cause primary ciliary dyskinesia with randomization of left/right body asymmetry. American journal of human genetics. 2013;93:357–367. doi: 10.1016/j.ajhg.2013.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Merveille AC, Davis EE, Becker-Heck A, Legendre M, Amirav I, Bataille G, Belmont J, Beydon N, Billen F, Clement A, Clercx C, Coste A, Crosbie R, de Blic J, Deleuze S, Duquesnoy P, Escalier D, Escudier E, Fliegauf M, Horvath J, Hill K, Jorissen M, Just J, Kispert A, Lathrop M, Loges NT, Marthin JK, Momozawa Y, Montantin G, Nielsen KG, Olbrich H, Papon JF, Rayet I, Roger G, Schmidts M, Tenreiro H, Towbin JA, Zelenika D, Zentgraf H, Georges M, Lequarre AS, Katsanis N, Omran H, Amselem S. CCDC39 is required for assembly of inner dynein arms and the dynein regulatory complex and for normal ciliary motility in humans and dogs. Nature genetics. 2011;43:72–78. doi: 10.1038/ng.726. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Zariwala MA, Gee HY, Kurkowiak M, Al-Mutairi DA, Leigh MW, Hurd TW, Hjeij R, Dell SD, Chaki M, Dougherty GW, Adan M, Spear PC, Esteve-Rudd J, Loges NT, Rosenfeld M, Diaz KA, Olbrich H, Wolf WE, Sheridan E, Batten TF, Halbritter J, Porath JD, Kohl S, Lovric S, Hwang DY, Pittman JE, Burns KA, Ferkol TW, Sagel SD, Olivier KN, Morgan LC, Werner C, Raidt J, Pennekamp P, Sun Z, Zhou W, Airik R, Natarajan S, Allen SJ, Amirav I, Wieczorek D, Landwehr K, Nielsen K, Schwerk N, Sertic J, Kohler G, Washburn J, Levy S, Fan S, Koerner-Rettberg C, Amselem S, Williams DS, Mitchell BJ, Drummond IA, Otto EA, Omran H, Knowles MR, Hildebrandt F. ZMYND10 is mutated in primary ciliary dyskinesia and interacts with LRRC6. American journal of human genetics. 2013;93:336–345. doi: 10.1016/j.ajhg.2013.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Wan J, Yourshaw M, Mamsa H, Rudnik-Schoneborn S, Menezes MP, Hong JE, Leong DW, Senderek J, Salman MS, Chitayat D, Seeman P, von Moers A, Graul-Neumann L, Kornberg AJ, Castro-Gago M, Sobrido MJ, Sanefuji M, Shieh PB, Salamon N, Kim RC, Vinters HV, Chen Z, Zerres K, Ryan MM, Nelson SF, Jen JC. Mutations in the RNA exosome component gene EXOSC3 cause pontocerebellar hypoplasia and spinal motor neuron degeneration. Nature genetics. 2012;44:704–708. doi: 10.1038/ng.2254. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Albers CA, Cvejic A, Favier R, Bouwmans EE, Alessi MC, Bertone P, Jordan G, Kettleborough RN, Kiddle G, Kostadima M, Read RJ, Sipos B, Sivapalaratnam S, Smethurst PA, Stephens J, Voss K, Nurden A, Rendon A, Nurden P, Ouwehand WH. Exome sequencing identifies NBEAL2 as the causative gene for gray platelet syndrome. Nature genetics. 2011;43:735–737. doi: 10.1038/ng.885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Norton N, Li D, Rieder MJ, Siegfried JD, Rampersaud E, Zuchner S, Mangos S, Gonzalez-Quintana J, Wang L, McGee S, Reiser J, Martin E, Nickerson DA, Hershberger RE. Genome-wide studies of copy number variation and exome sequencing identify rare variants in BAG3 as a cause of dilated cardiomyopathy. American journal of human genetics. 2011;88:273–282. doi: 10.1016/j.ajhg.2011.01.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Sanna-Cherchi S, Sampogna RV, Papeta N, Burgess KE, Nees SN, Perry BJ, Choi M, Bodria M, Liu Y, Weng PL, Lozanovski VJ, Verbitsky M, Lugani F, Sterken R, Paragas N, Caridi G, Carrea A, Dagnino M, Materna-Kiryluk A, Santamaria G, Murtas C, Ristoska-Bojkovska N, Izzi C, Kacak N, Bianco B, Giberti S, Gigante M, Piaggio G, Gesualdo L, Kosuljandic Vukic D, Vukojevic K, Saraga-Babic M, Saraga M, Gucev Z, Allegri L, Latos-Bielenska A, Casu D, State M, Scolari F, Ravazzolo R, Kiryluk K, Al-Awqati Q, D'Agati VD, Drummond IA, Tasic V, Lifton RP, Ghiggeri GM, Gharavi AG. Mutations in DSTYK and dominant urinary tract malformations. The New England journal of medicine. 2013;369:621–629. doi: 10.1056/NEJMoa1214479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Sarparanta J, Jonson PH, Golzio C, Sandell S, Luque H, Screen M, McDonald K, Stajich JM, Mahjneh I, Vihola A, Raheem O, Penttila S, Lehtinen S, Huovinen S, Palmio J, Tasca G, Ricci E, Hackman P, Hauser M, Katsanis N, Udd B. Mutations affecting the cytoplasmic functions of the co-chaperone DNAJB6 cause limb-girdle muscular dystrophy. Nature genetics. 2012;44:450–455. doi: 10.1038/ng.1103. S451–452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Veltman JA, Brunner HG. De novo mutations in human genetic disease. Nature reviews. Genetics. 2012;13:565–575. doi: 10.1038/nrg3241. [DOI] [PubMed] [Google Scholar]
- 70.Ramachandran KV, Hennessey JA, Barnett AS, Yin X, Stadt HA, Foster E, Shah RA, Yazawa M, Dolmetsch RE, Kirby ML, Pitt GS. Calcium influx through L-type CaV1.2 Ca2+ channels regulates mandibular development. The Journal of clinical investigation. 2013;123:1638–1646. doi: 10.1172/JCI66903. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Inoue K, Lupski JR. Molecular mechanisms for genomic disorders. Annual Review of Genomics and Human Genetics. 2002;3:199–242. doi: 10.1146/annurev.genom.3.032802.120023. [DOI] [PubMed] [Google Scholar]
- 72.Stankiewicz P, Lupski JR. Structural variation in the human genome and its role in disease. Annual review of medicine. 2010;61:437–455. doi: 10.1146/annurev-med-100708-204735. [DOI] [PubMed] [Google Scholar]
- 73.Lindsay EA. Chromosomal microdeletions: dissecting del22q11 syndrome. Nature reviews. Genetics. 2001;2:858–868. doi: 10.1038/35098574. [DOI] [PubMed] [Google Scholar]
- 74.Golzio C, Willer J, Talkowski ME, Oh EC, Taniguchi Y, Jacquemont S, Reymond A, Sun M, Sawa A, Gusella JF, Kamiya A, Beckmann JS, Katsanis N. KCTD13 is a major driver of mirrored neuroanatomical phenotypes of the 16p11.2 copy number variant. Nature. 2012;485:363–367. doi: 10.1038/nature11091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Dauber A, Golzio C, Guenot C, Jodelka FM, Kibaek M, Kjaergaard S, Leheup B, Martinet D, Nowaczyk MJ, Rosenfeld JA, Zeesman S, Zunich J, Beckmann JS, Hirschhorn JN, Hastings ML, Jacquemont S, Katsanis N. SCRIB and PUF60 are primary drivers of the multisystemic phenotypes of the 8q24.3 copy-number variant. American journal of human genetics. 2013;93:798–811. doi: 10.1016/j.ajhg.2013.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Davis EE, Katsanis N. The ciliopathies: a transitional model into systems biology of human genetic disease. Current opinion in genetics & development. 2012;22:290–303. doi: 10.1016/j.gde.2012.04.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Zaghloul NA, Liu Y, Gerdes JM, Gascue C, Oh EC, Leitch CC, Bromberg Y, Binkley J, Leibel RL, Sidow A, Badano JL, Katsanis N. Functional analyses of variants reveal a significant role for dominant negative and common alleles in oligogenic Bardet-Biedl syndrome. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:10602–10607. doi: 10.1073/pnas.1000219107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Khanna H, Davis EE, Murga-Zamalloa CA, Estrada-Cuzcano A, Lopez I, den Hollander AI, Zonneveld MN, Othman MI, Waseem N, Chakarova CF, Maubaret C, Diaz-Font A, MacDonald I, Muzny DM, Wheeler DA, Morgan M, Lewis LR, Logan CV, Tan PL, Beer MA, Inglehearn CF, Lewis RA, Jacobson SG, Bergmann C, Beales PL, Attie-Bitach T, Johnson CA, Otto EA, Bhattacharya SS, Hildebrandt F, Gibbs RA, Koenekoop RK, Swaroop A, Katsanis N. A common allele in RPGRIP1L is a modifier of retinal degeneration in ciliopathies. Nature genetics. 2009;41:739–745. doi: 10.1038/ng.366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.de Pontual L, Zaghloul NA, Thomas S, Davis EE, McGaughey DM, Dollfus H, Baumann C, Bessling SL, Babarit C, Pelet A, Gascue C, Beales P, Munnich A, Lyonnet S, Etchevers H, Attie-Bitach T, Badano JL, McCallion AS, Katsanis N, Amiel J. Epistasis between RET and BBS mutations modulates enteric innervation and causes syndromic Hirschsprung disease. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:13921–13926. doi: 10.1073/pnas.0901219106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Davis EE, Zhang Q, Liu Q, Diplas BH, Davey LM, Hartley J, Stoetzel C, Szymanska K, Ramaswami G, Logan CV, Muzny DM, Young AC, Wheeler DA, Cruz P, Morgan M, Lewis LR, Cherukuri P, Maskeri B, Hansen NF, Mullikin JC, Blakesley RW, Bouffard GG, Program NCS, Gyapay G, Rieger S, Tonshoff B, Kern I, Soliman NA, Neuhaus TJ, Swoboda KJ, Kayserili H, Gallagher TE, Lewis RA, Bergmann C, Otto EA, Saunier S, Scambler PJ, Beales PL, Gleeson JG, Maher ER, Attie-Bitach T, Dollfus H, Johnson CA, Green ED, Gibbs RA, Hildebrandt F, Pierce EA, Katsanis N. TTC21B contributes both causal and modifying alleles across the ciliopathy spectrum. Nature genetics. 2011;43:189–196. doi: 10.1038/ng.756. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Oprea GE, Krober S, McWhorter ML, Rossoll W, Muller S, Krawczak M, Bassell GJ, Beattie CE, Wirth B. Plastin 3 is a protective modifier of autosomal recessive spinal muscular atrophy. Science. 2008;320:524–527. doi: 10.1126/science.1155085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.van de Ven JP, Nilsson SC, Tan PL, Buitendijk GH, Ristau T, Mohlin FC, Nabuurs SB, Schoenmaker-Koller FE, Smailhodzic D, Campochiaro PA, Zack DJ, Duvvari MR, Bakker B, Paun CC, Boon CJ, Uitterlinden AG, Liakopoulos S, Klevering BJ, Fauser S, Daha MR, Katsanis N, Klaver CC, Blom AM, Hoyng CB, den Hollander AI. A functional variant in the CFI gene confers a high risk of age-related macular degeneration. Nature genetics. 2013;45:813–817. doi: 10.1038/ng.2640. [DOI] [PubMed] [Google Scholar]
- 83.Liu CT, Garnaas MK, Tin A, Kottgen A, Franceschini N, Peralta CA, de Boer IH, Lu X, Atkinson E, Ding J, Nalls M, Shriner D, Coresh J, Kutlar A, Bibbins-Domingo K, Siscovick D, Akylbekova E, Wyatt S, Astor B, Mychaleckjy J, Li M, Reilly MP, Townsend RR, Adeyemo A, Zonderman AB, de Andrade M, Turner ST, Mosley TH, Harris TB, Consortium CK, Rotimi CN, Liu Y, Kardia SL, Evans MK, Shlipak MG, Kramer H, Flessner MF, Dreisbach AW, Goessling W, Cupples LA, Kao WL, Fox CS. Genetic association for renal traits among participants of African ancestry reveals new loci for renal function. PLoS genetics. 2011;7:e1002264. doi: 10.1371/journal.pgen.1002264. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Gieger C, Radhakrishnan A, Cvejic A, Tang W, Porcu E, Pistis G, Serbanovic-Canic J, Elling U, Goodall AH, Labrune Y, Lopez LM, Magi R, Meacham S, Okada Y, Pirastu N, Sorice R, Teumer A, Voss K, Zhang W, Ramirez-Solis R, Bis JC, Ellinghaus D, Gogele M, Hottenga JJ, Langenberg C, Kovacs P, O'Reilly PF, Shin SY, Esko T, Hartiala J, Kanoni S, Murgia F, Parsa A, Stephens J, van der Harst P, Ellen van der Schoot C, Allayee H, Attwood A, Balkau B, Bastardot F, Basu S, Baumeister SE, Biino G, Bomba L, Bonnefond A, Cambien F, Chambers JC, Cucca F, D'Adamo P, Davies G, de Boer RA, de Geus EJ, Doring A, Elliott P, Erdmann J, Evans DM, Falchi M, Feng W, Folsom AR, Frazer IH, Gibson QD, Glazer NL, Hammond C, Hartikainen AL, Heckbert SR, Hengstenberg C, Hersch M, Illig T, Loos RJ, Jolley J, Khaw KT, Kuhnel B, Kyrtsonis MC, Lagou V, Lloyd-Jones H, Lumley T, Mangino M, Maschio A, Mateo Leach I, McKnight B, Memari Y, Mitchell BD, Montgomery GW, Nakamura Y, Nauck M, Navis G, Nothlings U, Nolte IM, Porteous DJ, Pouta A, Pramstaller PP, Pullat J, Ring SM, Rotter JI, Ruggiero D, Ruokonen A, Sala C, Samani NJ, Sambrook J, Schlessinger D, Schreiber S, Schunkert H, Scott J, Smith NL, Snieder H, Starr JM, Stumvoll M, Takahashi A, Tang WH, Taylor K, Tenesa A, Lay Thein S, Tonjes A, Uda M, Ulivi S, van Veldhuisen DJ, Visscher PM, Volker U, Wichmann HE, Wiggins KL, Willemsen G, Yang TP, Hua Zhao J, Zitting P, Bradley JR, Dedoussis GV, Gasparini P, Hazen SL, Metspalu A, Pirastu M, Shuldiner AR, Joost van Pelt L, Zwaginga JJ, Boomsma DI, Deary IJ, Franke A, Froguel P, Ganesh SK, Jarvelin MR, Martin NG, Meisinger C, Psaty BM, Spector TD, Wareham NJ, Akkerman JW, Ciullo M, Deloukas P, Greinacher A, Jupe S, Kamatani N, Khadake J, Kooner JS, Penninger J, Prokopenko I, Stemple D, Toniolo D, Wernisch L, Sanna S, Hicks AA, Rendon A, Ferreira MA, Ouwehand WH, Soranzo N. New gene functions in megakaryopoiesis and platelet formation. Nature. 2011;480:201–208. doi: 10.1038/nature10659. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Best JDA, K W. Zebrafish: An in vivo model for the study of neurological diseases. Neuropsychiatric Disease and Treatment. 2008;4:567–576. doi: 10.2147/ndt.s2056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Bandmann O, Burton EA. Genetic zebrafish models of neurodegenerative diseases. Neurobiology of Disease. 2010;40:58–65. doi: 10.1016/j.nbd.2010.05.017. [DOI] [PubMed] [Google Scholar]
- 87.Bai Q, Garver JA, Hukriede NA, Burton EA. Generation of a transgenic zebrafish model of Tauopathy using a novel promoter element derived from the zebrafish eno2 gene. Nucleic Acids Research. 2007;35:6501–6516. doi: 10.1093/nar/gkm608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Arthur D, Levin E. Spatial and non-spatial visual discrimination learning in zebrafish (Danio rerio) Anim.Cogn. 2001;4:125–131. [Google Scholar]
- 89.Wullimann MF, Mueller T. Teleostean and mammalian forebrains contrasted: Evidence from genes to behavior. The Journal of Comparative Neurology. 2004;475:143–162. doi: 10.1002/cne.20183. [DOI] [PubMed] [Google Scholar]
- 90.Auer TO, Duroure K, De Cian A, Concordet JP, Del Bene F. Highly efficient CRISPR/Cas9-mediated knock-in in zebrafish by homology-independent DNA repair. Genome research. 2014;24:142–153. doi: 10.1101/gr.161638.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Maddison LA, Li M, Chen W. Conditional gene-trap mutagenesis in zebrafish. Methods in molecular biology. 2014;1101:393–411. doi: 10.1007/978-1-62703-721-1_19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Jao LE, Wente SR, Chen W. Efficient multiplex biallelic zebrafish genome editing using a CRISPR nuclease system. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:13904–13909. doi: 10.1073/pnas.1308335110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Pardo-Martin C, Allalou A, Medina J, Eimon PM, Wählby C, Fatih Yanik M. High-throughput hyperdimensional vertebrate phenotyping. Nat Commun. 2013;4:1467. doi: 10.1038/ncomms2475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Kawakami K, Shima A, Kawakami N. Identification of a functional transposase of the Tol2 element, an Ac-like element from the Japanese medaka fish, and its transposition in the zebrafish germ lineage. Proceedings of the National Academy of Sciences of the United States of America. 2000;97:11403–11408. doi: 10.1073/pnas.97.21.11403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.McClellan J, King MC. Genetic heterogeneity in human disease. Cell. 2010;141:210–217. doi: 10.1016/j.cell.2010.03.032. [DOI] [PubMed] [Google Scholar]
- 96.Ragvin A, Moro E, Fredman D, Navratilova P, Drivenes Ø, Engström PG, Alonso ME, Mustienes EdlC, Skarmeta JLG, Tavares MJ, Casares F, Manzanares M, van Heyningen V, Molven A, Njølstad PR, Argenton F, Lenhard B, Becker TS. Long-range gene regulation links genomic type 2 diabetes and obesity risk regions to HHEX, SOX4, and IRX3. Proceedings of the National Academy of Sciences. 2010;107:775–780. doi: 10.1073/pnas.0911591107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Tan JL, Zon LI. Chemical screening in zebrafish for novel biological and therapeutic discovery. Methods in cell biology. 2011;105:493–516. doi: 10.1016/B978-0-12-381320-6.00021-7. [DOI] [PubMed] [Google Scholar]
- 98.Zon LI, Peterson RT. In vivo drug discovery in the zebrafish. Nature reviews. Drug discovery. 2005;4:35–44. doi: 10.1038/nrd1606. [DOI] [PubMed] [Google Scholar]