Next Article in Journal
Immunosuppressive Polymeric Nanoparticles Targeting Dendritic Cells Alleviate Lupus Disease in Fcgr2b-/- Mice by Mediating Antigen-Specific Immune Tolerance
Previous Article in Journal
Recent Advances in Characterization of Melanin Pigments in Biological Samples
Previous Article in Special Issue
Monitoring Nrf2/ARE Pathway Activity with a New Zebrafish Reporter System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modeling Human Muscular Dystrophies in Zebrafish: Mutant Lines, Transgenic Fluorescent Biosensors, and Phenotyping Assays

1
Department of Biotechnology, University of Verona, 37134 Verona, Italy
2
Department of Molecular and Translational Medicine, University of Brescia, 25123 Brescia, Italy
3
Neuroscience Institute, Italian National Research Council (CNR), 35131 Padua, Italy
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(9), 8314; https://doi.org/10.3390/ijms24098314
Submission received: 15 March 2023 / Revised: 28 April 2023 / Accepted: 3 May 2023 / Published: 5 May 2023
(This article belongs to the Special Issue Zebrafish as a Model in Human Disease)

Abstract

:
Muscular dystrophies (MDs) are a heterogeneous group of myopathies characterized by progressive muscle weakness leading to death from heart or respiratory failure. MDs are caused by mutations in genes involved in both the development and organization of muscle fibers. Several animal models harboring mutations in MD-associated genes have been developed so far. Together with rodents, the zebrafish is one of the most popular animal models used to reproduce MDs because of the high level of sequence homology with the human genome and its genetic manipulability. This review describes the most important zebrafish mutant models of MD and the most advanced tools used to generate and characterize all these valuable transgenic lines. Zebrafish models of MDs have been generated by introducing mutations to muscle-specific genes with different genetic techniques, such as (i) N-ethyl-N-nitrosourea (ENU) treatment, (ii) the injection of specific morpholino, (iii) tol2-based transgenesis, (iv) TALEN, (v) and CRISPR/Cas9 technology. All these models are extensively used either to study muscle development and function or understand the pathogenetic mechanisms of MDs. Several tools have also been developed to characterize these zebrafish models by checking (i) motor behavior, (ii) muscle fiber structure, (iii) oxidative stress, and (iv) mitochondrial function and dynamics. Further, living biosensor models, based on the expression of fluorescent reporter proteins under the control of muscle-specific promoters or responsive elements, have been revealed to be powerful tools to follow molecular dynamics at the level of a single muscle fiber. Thus, zebrafish models of MDs can also be a powerful tool to search for new drugs or gene therapies able to block or slow down disease progression.

1. Introduction

Muscular dystrophies (MDs) are a heterogeneous group of incurable hereditary diseases characterized by progressive weakness and the degeneration of skeletal muscle, leading to loss of ambulation, difficulties in the satisfaction of vital needs (i.e., breathing and eating), and death due to heart or respiratory failure [1].
Historically, MDs have been grouped according to the age of onset, the clinical and pathological manifestations, the pattern of inheritance (i.e., dominant, recessive, or sporadic due to de novo mutations), or the genetic causes as soon as they became clear [2]. MD subtypes such as laminopathies, sarcoglycanopathies, and dystroglycanopathies are, therefore, identified to underline the putative defective gene. However, such a classification, which is one of the most used, may result in some overlapping. Indeed, the same allelic mutations may give rise to a spectrum of clinical manifestations, and more than one gene may cause the same MD phenotype [2]. To further complicate the situation, the mutation of some MD-correlated genes has also been linked to diseases that do not primarily affect skeletal muscle [3,4,5]. In this review, the tremendous range of MDs was alternatively organized according to the localization/role of the mutant proteins that, over time, have been found to be disease-related. For example, the structural proteins of the basement membrane and extracellular matrix are distinguished from cytosolic proteins as well as the proteins involved in ER, and Golgi apparatus trafficking will be discussed separately from proteins involved in post-translational modifications.
Progress in understanding the pathogenetic mechanisms of MDs has been achieved thanks to several well-established experimental models. In this review, we provide a complete overview of the MD animal models generated in zebrafish (Danio rerio) so far.
Since the early 1980s, the zebrafish has emerged as a particularly valuable model organism [6,7] because of its unique advantages: (i) a small size and short generation time (adulthood is achieved in 3 months), (ii) the ease of husbandry and breeding, (iii) its high spawning productivity (up to 300 eggs once a week from a single mating event), (iv) its rapid and ex utero development, (v) and the optical clarity of the embryos and early larvae [8], as well as its genetic manipulability (see Section 4). The coexistence of the latter features makes this vertebrate organism particularly useful in the investigation of muscle functioning. In zebrafish, the skeletal muscle, which constitutes the main fraction of the body, becomes functional by the end of the first day of development, and it is fully differentiated after 48 h post-fertilization (hpf). The introduction of fluorescent transgenic tags may help to follow muscle development in living embryos via time-lapse confocal microscopy in a non-invasive manner [9]. Furthermore, as described in Section 1.1, zebrafish and human skeletal muscle show similarities in myofibrillar structure, contractile properties, and types of muscle fibers (fast and slow). Of course, the two species are evolutionarily far from each other, and they differ in several physiological and anatomical characteristics. For example, fast and slow muscle fibers in zebrafish are topographically separated, while they are mixed in humans. Additional disadvantages are the polyploidy of zebrafish for specific genes and a still limited number of cross-reacting antibodies. Despite these intrinsic limitations, the large amount of progress made over the last 20 years has demonstrated the usefulness of the zebrafish model in studying different aspects of human muscle diseases.

1.1. Skeletal Muscle Organization in Zebrafish

MDs are characterized by the progressive degeneration of skeletal muscle caused by gene mutations affecting muscle fiber organization and function. Some of them have been studied in experimental zebrafish models. Before going into the details of zebrafish MD models, a brief description of the skeletal muscle structure is necessary to clarify the role of some proteins mentioned in the following sections. The correct organization of the skeletal muscle plays a critical role in both humans and zebrafish, as it serves many purposes. It allows for movement through the contractile system, sustains the body’s architecture, maintains body temperature, stores nutrients, and stabilizes joints. In humans, skeletal muscle is a highly organized tissue composed of bundles of fascicles formed by groups of muscle fibers that contain several myofibrils composed of actin and myosin proteins. Each myofiber represents a unique multinucleated muscle cell along with its functional unit, the sarcomere. Myofibers are closed by the sarcolemma, which has a polysaccharide coating that is fused to tendon fibers that anchor myofibers to the tip of the myotendinous junction [10]. In zebrafish, skeletal muscle fibers are also organized into sarcomeres and myofibrils, but they are segmented by myosepta throughout the body into myomeres, which are morphological functional units [11,12,13] (Figure 1). The myoseptum provides structural support to the zebrafish skeletal muscle and facilitates the rapid contractions necessary for zebrafish swimming [14]. Teleost fish possess two distinct types of muscle fibers, which include slow-red and fast-white muscles (Figure 1). The slow muscle fibers can typically be identified in the narrow V-shaped area near the lateral line and are generally used for long and continuous swimming activities. These fibers are characterized by a relatively small diameter, aerobic metabolism, and an abundance of mitochondria and glycogen. Fast muscle fibers constitute the majority of skeletal muscle in teleost fish, and their fiber diameter is comparatively larger than the slow ones. These fast-twitching fibers are primarily utilized for high-speed swimming activities such as predation and escape [15]. The muscle sarcomere is the basic unit of muscle contraction and is responsible for the generation of force and movement. The organization of the sarcomere is highly complex, and it is regulated by a large number of proteins that are essential for its proper function. The main proteins responsible for force generation during muscle activity are myosin, actin, troponin, and tropomyosin. Myosin is a motor protein responsible for muscle contraction. It is the most abundant protein of the muscle myofibrils and, with actin, forms the basic contractile unit known as the actomyosin complex. Another major protein in muscle fibers is tropomyosin (TPM), a protein that regulates the interaction between actin and myosin, thus controlling muscle contraction [16]. In zebrafish, there are six tpm genes that include the human paralogs of the TPM1 (tpm1-1 and tpm1-2), the paralogs of the TPM4 gene (tpm4-1 and tpm4-2), and the two single-copy genes TPM2 and TPM3 [17]. Another structure that plays a major role in maintaining sarcolemmal integrity and function is the dystrophin-associated glycoprotein complex (DGC) (Figure 1). This protein structure connects the cytoskeleton to the extracellular matrix (ECM) and is composed of numerous cytoplasmic proteins such as dystrophin, syntrophin, and dystrobrevin, as well as transmembrane proteins such as dystroglycan, sarcoglycan, and sarcospan [18] (Figure 1). Outside the myofiber, laminin-2 binds to the DGC through α-dystroglycan, thus linking the sarcolemma to the ECM (Figure 1). Similarly, on the cytoplasmic side, dystrophin binds to F-actin and β-dystroglycan (Figure 1). In turn, α- and β-dystroglycans, which are closely associated, form a stable membrane complex, providing additional structural support to the DGC that is essential for transmitting force and stabilizing the membrane during contraction [19]. Losing DGC structure results in the disruption of sarcolemma upon mechanical stress, and it is not surprising that mutations occurring in DGC-associated genes are often linked to the onset of severe forms of MD. Mutations in the DMD gene, which encodes dystrophin, cause Duchenne muscular dystrophy (DMD) and Becker muscular dystrophy (BMD), one of the most studied forms of MD. Moreover, mutations in genes encoding for dystroglycan, such as DAG1 or Fukutin-related protein (FKRP), cause various forms of MD, including muscular dystrophy–dystroglycanopathy type A9 (MDDGA9) or limb–girdle muscular dystrophy (LGMD) type 2I [20,21,22]. Mutations occurring in the genes that encode sarcoglycans can be associated with different forms of LGMD, such as LGMD2C, LGMD2D, LGMD2E, and LGMD2F [23]. The loss of sarcoglycans destabilizes the dystrophin complex, increasing susceptibility to mechanical stress and damage to muscle fibers. Finally, in addition to the DGC core proteins, the disruption of other molecules can be also associated with MDs. These proteins include the matrix protein laminin A2, which is linked to congenital MDs [24]; the protein telethonin, associated with LGMD2G [25]; or the transmembrane protein dysferlin, which plays a role in muscle membrane repair and regeneration. Mutations in the DYSF gene, which encodes dysferlin, can cause several types of MD, including LGMD2B [26]. Moreover, extracellular proteins such as collagen VI, the calcium-activated protease calpain, and the intermediate filament protein lamin A/C, have also been associated with the onset of Ullrich/Bethlem congenital muscular dystrophy [27], LGMD2A [28], and Emerye–Dreifuss/LGMD1B dystrophy, respectively [29]. In conclusion, the proteins involved in muscular dystrophy play essential roles in maintaining the structure and function of muscle cells. Mutations in genes encoding these proteins lead to the loss of stability and integrity in the muscle cell membrane, making the muscle fibers more susceptible to damage and degeneration. Understanding both the molecular mechanisms underlying the total or partial absence of these proteins and their main interactions is critical to developing effective therapies for MDs.

2. Models of Muscular Dystrophy in Zebrafish

2.1. Basal Membrane and Extracellular Matrix Protein

The ECM comprises 1–10% of the total muscular weight, depending on the skeletal muscle, and it has several fundamental functions such as force transmission, support, and the repair of muscle fibers after damage. Moreover, the ECM provides a particular architecture that maintains myofibers, blood vessels, and nerves in close contact inside the muscle [30]. In addition, the composition of the ECM may influence the contractile properties and the gene expression of the muscular fibers, changing the rigidity of the basal membrane [31]. Due to all these characteristics, alteration in the composition and components of the ECM can lead to the onset of several pathologies, including MDs.
As described previously, one of the most important EMC components is laminin α2, which is encoded by LAMA2. In humans, mutations of this gene can give rise to the development of a range of MD spectra, from the more frequent and aggressive Merosin-deficient congenital muscular dystrophy 1A (MDC1A) to a milder pathology reassembling the LGMD phenotype with later-childhood onset [32]. The dystrophic zebrafish mutant candyfloss (cafteg15a), which is generated by N-ethyl-N-nitrosourea (ENU) mutagenesis, shows non-sense mutations in the globular domain of the laminin α2, mimicking the clinical genetic lesions of MDC1A patients [8]. Interestingly, candyfloss mutants display muscle fiber detachment from the myotendinous junction (MTJ) starting by 72 hpf, with muscle degeneration and, consequently, altered motility as well as fewer coils [8,33]. In addition, homozygous mutants in most cases present death by 16 days post-fertilization (dpf) and a very small or absent production of progeny [8]. Moreover, muscle fiber detachment is induced by motor activity [8], in which the mechanical overload of mutant muscles increases the severity and progression of the pathology, though embryo immobilization can salvage the disease phenotype [34]. However, unlike other MD models, such as sapje (dmdtm90c) fish, where the dystrophic fibers are short-lived due to sarcolemma disruption, in the candyfloss mutant, the muscle fiber detachment does not affect sarcolemma integrity, and the myofibers live longer [35,36]. The candyfloss muscle fibers show the ability to undergo cellular remodeling, in particular forming membrane protrusions and branches [35]. Moreover, after detachment, the myofibers go through a process of fusion, thus demonstrating that several survival mechanisms are activated in post-detachment muscle fibers. Importantly, such behavior seems to be specific to LAMA2 dystrophic fibers, since dystrophin mutant myofibers do not display any of these features [35]. Furthermore, candyfloss myofibers present an abnormal lysosome distribution at the sarcolemma, in which these organelles can be involved in mechanisms of membrane repair and the removal of cellular debris after detachment, though further evidence is needed to understand the specific function of this phenomenon [37]. Autophagy does not appear to be the only altered cellular process, considering that the proteasome degradation pathway seems to be upregulated in different types of MD, including MDC1A. Overexpression of the proteasome complex and an increase in protein ubiquitination can be used as dystrophies hallmarks and inhibition of proteasome activity with the MG-132 inhibitor can rescue the pathological phenotype [37].
Candyfloss is not the only zebrafish model for the diseases related to LAMA2 mutations, but another zebrafish line, called lama2cl501, has been identified through ENU mutagenesis and further analyzed. The lama2cl501 mutant contains a mutation in the coiled-coil domain in the long arm of LAMA2, thus resulting in the loss of interaction between laminin α2 and the other two laminins of the heterotrimeric complex [38]. The phenotype of this mutant overlaps with that of the candyfloss zebrafish, in which the main pathological manifestation is the detachment of myofibers from the MTJ without disrupting the sarcolemma. Moreover, the lama2cl501 mutant displays other phenotypic features, such as lower laminin α2 expression in the basal membrane, smaller myotomes, defects in the sarcomere architecture, and brain and eye alterations [38].
Although laminin α2 is one of the major components that maintain attached muscle fibers, other laminins may be involved in anchoring myofibers to the MTJ, since injecting the LAMA1 morpholino into LAMA2 mutants results in a more severe phenotype [39]. This is also suggested by the fact that the ilk/dmd and ilk/dag1 double-homozygous mutants present a more severe disease than the LAMA2/dmd and LAMA2/dag1 mutants. This also demonstrates that proteins directly or indirectly interacting with laminin α2, such as dystroglycan, dystrophin, and ilk, if altered, can modify the LAMA2-MD phenotype [39,40].
As outlined above, collagens also have different fundamental functions such as providing elasticity, support, and organization to the ECM. In particular, collagens are essential in forming the MTJ and play a major role in keeping myofibers attached to the ECM [41]. For this reason, the different zebrafish models generated with morpholine, TALEN, or CRISPR/Cas9 mutagenesis display similar features (see Table 1). Mutations in one of the three genes that codify the components of collagen VI lead to the onset of either Ullrich congenital MD (UCMD) or Bethlem myopathy (BM) [42], while alterations in the col22a1 gene can cause disruption in the MTJ [43]. In particular, the different zebrafish lines exhibit alterations in the organization of slow muscles and the ECM throughout the different stages of their life cycles. This phenomenon becomes more evident during mechanical stress, thus inducing the detachment of the myofibers from the MTJ, reassembling a phenotype similar to the candyfloss model [44]. Moreover, both the morphants and mutants show abnormal mitochondria, an abnormal endoplasmic reticulum (ER), defective autophagy, and fibrosis [44,45,46]. Interestingly, as the col6a1ama605003 mutant ages, it develops a stress behavior due to hypoxia, thus leading to several respiratory-adaptive responses such as residing in a surface area with a higher concentration of oxygen and increasing the respiratory rate [45].

2.2. Cytosolic Proteins

Dystrophin was the first protein identified as causative of MD. It localizes in the cytosol, although it is directly related to the muscle sarcolemma as a founder of the dystrophin-associated protein complex (DAPC). From its early discovery in the late 1980s, many mammalian and non-animal models have been generated. In 2002, Parsons et al. [48] described the first zebrafish model for skeletal MD, where the dystroglycan was knocked down by a morpholino oligonucleotide (MO), which created an MD phenotype. One year later, Bassett et al. [49] molecularly characterized the first stable zebrafish MD model, known as sapje. This mutant was earlier identified by Granato and colleagues [50] during a mobility dysfunction screening of zebrafish mutagens generated using ENU at Tübingen University. In particular, the mutation harbored by sapje led to reduced striation and somite degeneration; although the zebrafish embryos presented with muscle fiber formation comparable to the wild-type, at 96 hpf, the swimming velocity and muscle birefringence were reduced. During development, the lesions become severe in the somitic segments, and the degeneration was restricted to skeletal muscle tissue, but heart and jaw muscles were not affected. Basset et al. [49] described the degeneration mechanism of full sapje embryo mutants as the separation of skeletal muscle fibers from their attachment points on myosepta. Furthermore, they identified the sapje locus as dmd [51], nowadays well known as the Danio rerio ortholog of the Homo sapiens DMD gene, and this showed that the mutant carries a non-sense mutation in exon 4 of the zebrafish dystrophin gene.
The characterization of the zebrafish sapje mutation revealed a novel pathological mechanism together with the validation of a model for DMD and BMD that is currently not only a worldwide landmark but also a valuable platform for compound screenings, thus allowing for new insights into a wide range of diverse muscle diseases.
An additional zebrafish dystrophic mutant was also screened via the early-pressure technique and genetically isolated based on the symptoms of muscle disease presence by Guyon et al. [52]. They identified 8 out of 500 different mutants presenting abnormal birefringence, among which, one had a phenotype very similar to the sapje model, and it was thus called sapje-like. Like sapje zebrafish, the sapje-like model showed clear muscle degeneration at 5 dpf, and sequencing analysis showed that it carries a mutation in the donor splice junction of exon 62 of the dystrophin gene. These results represented the characterization of the first zebrafish model harboring a splice-site mutation in the dystrophin gene, exploitable for new studies and therapies for muscle disorders associated with splicing.
Another cytosolic protein involved in the dystrophy scenario is DUX4. Its role in late-onset facioscapulohumeral dystrophy type 1 (FSHD-1), which is the most common autosomal dominant form of MD, was described by Kunkel’s research team, who generated a dux4 injection model (DUX4i) [53] and a dux4 transgenic model (DUX4t) [54]. In particular, despite the FSHD-1 onset in the second decade of life, DUX4 misexpression triggers muscle degeneration in the early stages and, therefore, it appeared to have a development role rather than being causative. During the larval period (3–5 dpf), the two mutants presented abnormal birefringence and late hatching; the DUX4i also showed abnormal eye and ear formation, together with fin asymmetry, while DUX4t swam shorter distances and had a lower twitch and tetanic force. This phenotype was mild in adult DUX4i mutants, whose lipids and collagen were neutral together with inflammation and muscle function. They also showed fin and muscle regeneration. On the contrary, DUX4t mutants developed abnormal muscle, asymmetrical fat deposition, and collagen accumulation. Furthermore, muscle function was fully abnormal. The DUX4 developmental role was also confirmed by Wallace et al. [55], who over-expressed the gene in zebrafish embryos. At 4 dpf, the mutants showed defects in muscle histology, curved bodies, fin abnormalities, and cardiac hypertrophy.
Particularly noteworthy are the stable muscleblind-like (mbnl) zebrafish mutants generated by Hinman and colleagues to study the pathogenetic mechanisms of myotonic dystrophy [56]. The MBNL family is composed of three RNA-binding proteins (MBNL1, MBNL2, and MBNL3), which are highly conserved across several species. In zebrafish, loss-of-function mutations associated with these proteins lead to physical and molecular alterations consistent with those present in myotonic dystrophy patients with mutations in the DMPK and ZNF9 genes. Interestingly, single, double, and triple mbnl homozygous zebrafish showed a relevant disease phenotype, including impaired motor behavior and decreased body size [56]. Transient mbnl zebrafish models were also generated [57,58], but their characterization pointed out several limitations compared with the stable mutant models [56].
Moreover, two CRISPR/Cas9 zebrafish KO models for the Valosin-containing protein (vcp) gene were recently generated and validated for mechano-pathological studies of VSPopathies. VCP is a type II AAA ATPase involved in cellular protein homeostasis that has been associated with the onset of IBMPFD syndrome and some rare variants of ALS or myofibrillar myopathy. In two recent works, the 72 hpf old KO crispants of the gene showed (i) a reduction in the birefringence signal, (ii) disorganization in both myofibrils and sarcomeres, (iii) dysmorphic mitochondria, and (iv) a significant reduction in the distance moved after a touch-evoked escape response test [59,60]. The phenotype of the vcp crispants, also showing defects in cardiomyocyte functions and disruption in protein homeostasis, recapitulated skeletal muscle myopathy, which is consistent with previous findings and previous morpholino-mediated vcp knockdowns [59,60,61].
The disruption of many other cytosolic proteins has been associated with the onset of different dystrophy forms, but only transient zebrafish models have been described, as listed in the following Table 2.

2.3. Dystroglycan and α-DG Glycosylation-Related Protein

As described in Section 1.1, dystroglycan (DG) is an essential and conserved DAPC component that plays both a structural and functional role. It is encoded by the homonymous gene (DAG1) and consists of two subunits, the transmembrane β-DG and the extracellular α-DG, which requires complex post-translational modifications, particularly glycosylation, by a group of enzymes to exert its ligand-binding function [73].
DG impairment has been associated with a variety of MDs, generically known as dystroglycanopathies (DGPs). In particular, mutations of the glycoprotein DG have been grouped as primary DGPs, whereas secondary DGPs are caused by defective DG glycosylation. Nowadays, about 20 genes have been linked with secondary DGP [32]. Since their mutation is often associated with different phenotypes, it is becoming clearer that the disease severity is instead largely correlated with the level of α-DG hypo-glycosylation, and the upregulation/rescue of a single enzyme could be beneficial for multiple forms [74].
The first zebrafish model of DGP was described by Parsons and colleagues [48], who used an antisense morpholino oligonucleotide (MO) to knock down DG. The resulting morphant embryos were characterized by delayed development together with a hooked tail, less flexibility, uncoordinated movement, and compromised muscle integrity [48]. In zebrafish, the complete absence of both the α- and β-DG subunits was subsequently achieved via ENU mutagenesis screening [75]. The resulting model system, named the pathchytail (dag1cl500) mutant, was characterized by a late onset of the MD phenotype, with muscle birefringence defects becoming evident at 7 dpf, as well as disorganized T-tubules. Moreover, brain abnormalities and ocular defects have commonly been observed in pathchytail mutants, suggesting additional roles for DG in brain development [75].
As mentioned before, secondary DGPs are caused by defects in α-DG glycosylation, a modification necessary to properly perform its function. Over the years, several dystrophic-like zebrafish models have been generated by the morpholino-mediated knockdown of orthologous genes encoding for enzymes whose mutations are associated with secondary DGPs in mammals. The down-regulation of these proteins results in fish showing many defects, such as muscular disorganization, severe neurological and ophthalmic impairments, and often early death—for more details, see Table 3. Particularly noteworthy are stable fkrp zebrafish mutants, which were recently generated by using different genome editing strategies [76,77]. The fkrp zebrafish mutant, generated by Serafini et al. [76], phenocopies the symptoms observed in patients affected by Walker–Warburg syndrome, such as muscle breakdowns, head malformations, and early death. Thus, it could be a useful tool for future high-throughput drug screenings to identify new therapeutical compounds for MD treatments. Moreover, a second fkrp zebrafish mutant revealed, for the first time, that fkrp is also involved in fibronectin sialylation occurring in the Golgi, an essential process both for fibronectin–collagen binding and for the maintenance of the muscle basement membrane integrity [77]. Such observations may be a possible explanation for the wide spectrum of DGP clinical manifestations.

2.4. Nuclear Envelope and Cytoskeleton Proteins

The skeletal muscle cytoskeleton is organized in a complex way to coordinate muscle contractions. The fundamental unit is the sarcomere, composed of several structural and regulatory proteins, for which interaction with intermediate filaments is necessary both for the maintenance of cellular integrity and for the transduction of mechanical stimuli. Further, the intermediate filament lattice links the entire contractile apparatus to the sarcolemma and other organelles, such as the nucleus, mitochondria, lysosomes, and the sarcoplasmic reticulum [92]. The intermediate filament (IF) network is mainly composed of the muscle-specific IF protein desmin, which is associated with other IF proteins, such as synemin, paranemin, and syncoilin, as well as lower amounts of keratins K8 and K19 [92]. Desmin and its associated proteins form a network around the Z-disks, connecting them to each other and to the sarcolemma [93]. Except for paranemin, all the IF proteins are associated with costameres. This association is mediated by dystrobrevin (through synemin and syncoilin), dystrophin (through synemin and keratin 19), and myospryn. Keratin filaments (K8/K19) localize around the M-lines and near the cell surface, where they also link to costameres. The complex network of the IF lattice also provides effective functional crosstalk between organelles, such as the ER and mitochondria, and promotes lipid and calcium traffic. Thus, it is not surprising that alterations to the integrity of the IF network might lead to cardiac or skeletal myopathy [92].
Mutations in the desmin (DES) gene lead to myofibrillar myopathies, a group of muscle disorders characterized by the presence of heterogeneous desmin protein aggregates [94]. Paulin and colleagues provided novel insights into the link between aggregate deposition and myofibrillar degeneration [95] by generating two zebrafish knock-in lines named desmaCt122aGt and desmaCt122aRGt. These two mutant lines revealed both that the protein aggregation does not simply depend on specific mutations occurring in the desma gene and that protein aggregate deposition leads to myofibrillar alterations similar to those observed in the KO desmasa5−/− zebrafish mutant line, previously obtained via ENU mutagenesis screening. It also presented perturbed organization in the myofibrils, together with a decreased swimming capacity [94]. Similar skeletal muscle phenotypes were also observed in desma morphants, together with a reduction in the hatching rate, spontaneous movements, body length, time, and distance moved, as described in Li et al., 2013 [96] and Bührdel et al., 2015 [61]. On the contrary, knocking down the zebrafish desmin genes using CRISPR/Cas9 did not lead to skeletal muscle degeneration but altered calcium flux in myofibers [97].
Nicolas and colleagues [98] described the first transgenic zebrafish model for the laminin A/C gene, generated by using the CRISPR/Cas9 mutagenesis system. Zebrafish lmna mutants display alterations in skeletal muscle organization, transient cardiac abnormalities, and anomalous motor behavior; in fact, the mutants swim less and slower than normal, mimicking the Emery–Dreifuss muscular dystrophy (EDMD) phenotype. Indeed, both homozygous and heterozygous mutants show skeletal muscle damage, with an irregular distance between the myofibers and cytoskeletal modifications due to the disruption of F-actin, which can no longer interact with laminin-A/C, thus destabilizing the nucleo-cytoskeletal architecture and impairing the response to mechanical stimulus. Moreover, the upregulation of phosphorylated Pkc α and Erk 1-2, as well as changes in skeletal-muscle-gene-related expression (e.g., aberrant MAP kinase signaling pathway activation), were observed in both crispant and morphant zebrafish models for laminin-A/C [98].
Another interesting zebrafish model is the stretched-out (sotp1cpej) mutant, characterized by a genetic lesion of the flncb gene that is associated with the early onset of a severe form of myofibrillar myopathy. This mutant shows damaged or disrupted slow muscle fibers but no alteration in fast fibers at 26 hpf. Further analysis revealed that embryos at 32 or 48 hpf did not present any alteration, thus suggesting that defective fibers can go through turnover and can be replaced by new slow muscle fibers [99]. Indeed, the turnover and replacement of fibers in the sot mutant may be due to a change in gene expression, from flncb to flnca, at the level of the primary slow muscle fibers. This indicates that the two gene isoforms are functionally redundant and that the total level of Flnc expression is crucial to fiber maintenance. This is also demonstrated by the fact that the loss of flnca causes a mild phenotype very similar to that of the sot mutant, whereas the loss of both isoforms leads to weakness and the almost total loss of slow muscle fibers [100].
Other nuclear and cytoskeletal components are involved in the development of a wide spectrum of MDs, whose models are summarized in Table 4.

2.5. Membrane Proteins

Caveolin-3 is one of the major integral membrane proteins related to caveolae, small invaginations of the plasma membrane abundant in animal cells [106]. It is mainly expressed in cardiac and skeletal muscle cells, and mutations in the caveolin-3 gene (cav3) have been identified in human patients with different muscle diseases, including LGMD [107].
Several zebrafish models have been generated for cav3 using techniques such as mRNA injection, morpholinos, and the Tol2 transposon system [63,108]. The resulting mutants display large intercellular space and shorter mononucleated myofibrils, showing oval and ramified shapes as well. Moreover, morphants show disorganized contractile apparatuses, abnormal endomembrane systems, and the upregulation of eng1a gene expression, all of which induce an increase in muscle progenitors neighboring the notochord, thus demonstrating the main role of caveolin-3 in muscle maintenance and fiber fusion [108]. Interestingly, the transgenic cav3 dystrophic mutant shows membrane disruption but does not show evidence of fiber detachment like in the candyfloss and sapje mutants [8,49], suggesting another mechanism leading to membrane damage. The consequent weakness in the sarcolemma may trigger apoptosis, demonstrating the importance of caveolin-3 for membrane integrity, especially during muscular activity [63].
Furthermore, zebrafish morphant models have also been generated for other membrane proteins, resulting in skeletal muscle damage and disorganization as well as the abnormal development of the eyes and brain (see Table 5).

3. Functional and Genetic Tools for Investigating Skeletal Muscle Development and Functions

All previously described zebrafish models of myopathies are useful tools to study these disorders, as they share similar features with human diseases. Most of them can be easily characterized using many functional, behavioral, and phenotypical assays to test motor behavior, muscle force, muscle structure, calcium homeostasis, oxidative stress, mitochondrial function, and metabolism.

3.1. Motor Behavior

Several methods are available to assess motor behavior in zebrafish at different stages of development. Zebrafish embryos, larvae, and adults show different stereotyped motor behaviors depending on both morphological and functional changes in cells related to locomotion, such as muscle fibers, motor neurons, and cells involved in the development of the swim bladder.
At 17 hpf, the time when muscles start to be innervated by primary motor neurons, spontaneous coiling contractions are observed, which reach a peak of frequency within 3 h and progressively decline until 26 hpf [112].
By 27 hpf, embryos start to respond to mechanical stimuli, swimming around a Petri dish if touched with a little tip near the nerve centers [46]. More mature swimming, linked to beat-and-glide responses, appears around 5 dpf.
In this context, (i) spontaneous coiling events or tail flips (to assess side-to-side coiling contractions) [113], (ii) responses to external tactile stimuli (touch-evoked escape response test) [114], (iii) and swimming motor behavior (Noldus DanioVision System recording), have been extensively described and used through the years to evaluate dystrophic lines [46,115,116,117,118,119].
Spontaneous coiling contractions can be recorded with light microscopy as the number of events observed in 15 s or more for individual embryos.
The touch-evoked escape response test is performed by touching embryos and evaluating their ability to escape after a stimulus. This test is frequently measured using a scale of four values: 3, normal motor behavior; 2, mild motor impairment; 1, coiling events or flickers of movement both related to a severe motor impairment; 0, complete paralysis [46,119]. In addition, the measurement of these two assays can be automated [120].
Zebrafish motor behavior comprises a complex repertoire within a window between 17 and 120 hpf. It can be analyzed using different video tracking hardware and software, which provide readouts on zebrafish behavioral responses to a great variety of stimuli: light/dark, acoustical, environmental, and compounds [121,122,123].
As proposed by Stocco et al. [118], zebrafish larvae can be stimulated to swim using a time routine, repeated several times, involving exposure to 5 min of light periods after 30 min of incubation in the dark. After the test, the total distance moved can be analyzed.
Recently, a swimming tunnel instrument was developed to measure different parameters linked to the motor behavior of both zebrafish larvae and adults. It consists of a glass tube, filled with fish water, connected to a specific instrument that measures parameters, such as distance moved, muscle force, oxygen consumption rate, and metabolic changes, which will be recorded on a computer equipped with specific software [124,125,126,127,128].

3.2. Muscle Force (Electrophysiology)

Muscle force in zebrafish larvae should be analyzed also using electrophysiology. Briefly, larvae between 48 and 120 hpf are anesthetized and decapitated. Then, the body trunk is transferred to an experimental chamber, and the sample’s proximal end is coupled to a titanium wire connected to an isometric force transducer. The distal end of the titanium wire is linked to a high-speed motor lever arm. Muscles are then stimulated with a biphasic muscle stimulator through platinum electrodes [129,130].
The electrophysiological method has been successfully applied to different zebrafish models of MD, which show a progressive reduction in contractile force during disease progression [130,131,132,133].

3.3. Muscle Structure

The fastest and most reliable method to assess muscle fiber structure in zebrafish is birefringence, a technique that measures muscle anisotropy. which is the ability of muscle fibers to refract polarized light [50,134]. This technique offers the possibility of identifying and quantifying muscle damage in zebrafish at 40 hpf in a less invasive way [110,135]. Birefringence is also applied routinely to perform high-throughput drug screening and test drug efficacy [136,137].
The technology most used to observe the detailed ultrastructure of muscle tissue is transmission electron microscopy (TEM). This high-performance technique offers the possibility of obtaining qualitative and quantitative data concerning sarcomeric organization, the basement membrane, the MTJ, and the myoseptum structure, together with mitochondrial alterations, such as outer mitochondrial membrane detachment and swelling; quantitative analyses of distances between the outer and inner mitochondrial membrane can be also performed [11,43,117,118].
The availability of transgenic zebrafish biosensors [138] and several specific fluorescent antibodies help visualize the structure of a single muscle fiber using the most sophisticated lightsheet, confocal, and two-photon microscopy. In particular, antibodies (as described in the protocol of Gody and Henry, 2013 [139]) have been used to visualize filamentous actin and changes in the organization of myofibers in several zebrafish dystrophic muscle models [33,77].

3.4. Calcium Homeostasis and Oxidative Stress (Use of Fluorescent Probes)

Several methods based on genetically encoded biosensors that could be directly injected into zebrafish embryos or integrated into the zebrafish genome are used to visualize tissue-specific metabolic changes. Genetically encoded biosensors (FiNad sensors) of a wide dynamic range can easily be used to track cytosolic NAD+ and NADH redox states in vivo [140,141]. Other specific biosensors, such as Laconic [142], could be useful in assessing in vivo metabolic changes by observing lactate concentrations in muscles. Recently, two other genetically encoded biosensors, Rogfp2-Orp1 and Grx1-Rogfp2, were developed to measure in vivo hydrogen peroxide levels and glutathione redox potential, respectively [143]. Zebrafish transgenic lines that specifically express these biosensors in the muscle fibers could be another useful tool to explore the redox state in living animals. The measurement of mitochondrial reactive oxygen species (ROS) can be also achieved using fluorescent probes, such as the mitoSOX probe, which is able to react with ROS derived by both hydrogen peroxide and oxygen, as in the protocol of Rissone and Candotti [118,144,145].
The dysregulation of Ca2+ ions in muscle fibers is a primary cause of fiber death. Thus, different genetically encoded calcium indicators have been developed to check both calcium levels and dynamics in vivo at the level of a single muscle fiber [146,147,148]. These indicators are based on modifications of calmodulin (CaM), in which the calcium-binding domain is fused to GFP. GCaMP indicators with different calcium sensitivities have been developed (GCaMP1, GcaMP3, GcaMP6, GcaMP6f, GcaMP6s, and GcaMP7) to check the cytoplasmic levels of calcium [147,149,150]. Modified GCaMP6f, which harbors a mitochondrial-specific leader peptide at the N-terminus, was developed to check calcium levels in mitochondria [151].

3.5. Mitochondrial and Metabolic Function

The mitochondrial functions of zebrafish models of MD can be assessed with several methods.
The most common assay to evaluate the mitochondrial membrane potential is the accumulation of the tetramethylrhodamine methyl ester (TMRM), a fluorescent potentiometric probe targeting the mitochondrial membrane. This assay can be performed by incubating zebrafish embryos at 24/48 hpf with 300 nM of TMRM and 1.6 μM of CsH (to inhibit the multiple drug resistance pumps) in the dark for 24 h [118].
A powerful technique to measure (i) mitochondrial and cell respiration through the oxygen consumption rate (OCR) and (ii) metabolic changes through the extracellular acidification rate (ECAR) is the Seahorse (Agilent Technology), which has been successfully applied to zebrafish, as demonstrated by Bond et al., 2018 [152]. All measurements can be performed on a single zebrafish sample (between 24 and 96 hpf) placed in the well of an islet capture microplate and treated sequentially with oligomycin, FCCP, rotenone, and antimycin A [117,118].
Mass spectrometry possesses great potential to characterize the proteome of zebrafish models of MD to identify both mechanisms involved in disease pathogenesis and new potential druggable targets [153]. It also offers the possibility of evaluating the efficacy and toxicity of potential drugs [154] and obtaining specific profiles of protein modifications, such as the glycome profile [155]. Mass spectrometry can be performed at all stages of zebrafish development, from embryos to adults, by isolating proteins from entire animals or specific tissues.

4. Transgenic Fluorescent Zebrafish Lines

4.1. Zebrafish Reporter Lines Targeting In Vivo Signaling Pathway Activities Involved in Muscle Development

In this section, we provide a novel approach to system biology using transgenic biosensor animals that express fluorescent proteins under the control of signaling pathway-responsive cis-elements.
Pathway-specific responsive/regulatory elements (REs) can be multimerized to increase the responsiveness of the intended reporter and possibly attract more signal-dependent transcription factors. With a minimal promoter and a polyA signal site, multimerized REs are cloned upstream of a reporter gene encoding a fluorescent protein [138].
Bone morphogenetic proteins (BMPs), the wingless-related integration site (Wnt), sonic hedgehog (Shh), the hypoxia-inducible factor (HIF), the Nrf2-ARE pathway (Nrf2/ARE), the glucocorticoid reporter (GRE-reporter), and cAMP-response element-binding protein (CREB) signaling are crucial in controlling cell proliferation, stem cell maintenance, and differentiation in skeletal muscle. Several transgenic zebrafish lines have been created to investigate the aforementioned signaling pathways and the effects of their hypo- or hyperactivation in skeletal muscle (Table 6).
The HIF reporter zebrafish, named Tg(4xHRE-TATA: eGFP)ia21 [156], expresses tandem copies of a hypoxia response element (HRE), driving eGFP expression. The HIF pathway is active during zebrafish development, and transgenic embryos exhibit fluorescence from 24 hpf onward in muscle, lenses, skin, hearts, retinas, neurons, notochords, and the circulatory system.
The Shh signaling pathway has been demonstrated to be involved in skeletal muscle development in zebrafish [157]. To visualize in vivo Shh signaling with high detection sensitivity and a single-cell resolution, different transgenic zebrafish were generated carrying Gli-dependent fluorescent reporters. The Shh-pathway-driven expression of exogenous genes in cultured cells or live organisms was previously achieved using a minimal δ-crystallin promoter and eight tandem Gli binding sites derived from the murine Fox2A floor plate enhancer [158]. Mich et al. prepared reporter constructs that coupled these regulatory elements with sequences encoding nuclear-localized fluorescent proteins tagged with the ornithine decarboxylase-derived destabilizing peptide Tg(8xGliBS:mCherry-NLS-Odc1)st1002. Moreover, Corallo et al. generated transgenic lines that coupled these regulatory elements with sequences encoding localized fluorescent protein alone (eGFP or mCherry). The Shh reporter line shows fluorescence activity in the somitic region of a Tg(8xGliBS:mCherry-NLS-Odc1)st1002 zebrafish since 24 hpf, allowing for the labeling of muscle progenitor cells and superficial slow-twitch muscle fibers [159,160].
Wnt/β-catenin signaling in live embryos can be studied using several transgenic lines. The fluorescent reporter in the Tg(top:GFP)w25 transgenic line is under the control of TOPFlash, which consists of four consensus TCF/LEF binding sites placed next to a c-fos minimal promoter [161,162]. Additionally, transgenic fluorescent zebrafish lines, called Tg(7xTCF-Xla.Siam:GFP)ia4 and Tg(7xTCF-Xla.Siam:nlsmCherry)ia5, which are driven by the enhancer region of the minimal promoter of the Xenopus gene siamois, with seven tandem repeats of the TCF/LEF binding elements, have been developed [163]. With these reporter lines, Moro et al. [163] identified zebrafish tissues sensitive to canonical Wnt stimuli, including larval brains, notochords, otic vesicles, somites, fin buds, eyes, and posterior lateral lines.
Pyati et al. [164] created a transgenic zebrafish line that had a dominant negative BMP receptor linked to GFP integrated into its genome under the control of the heat shock promoter Tg(hsp70l:dnBmpr-GFP). By subjecting this transgenic line to heat shock at various embryonic stages, the researchers were able to distinguish between the functions of BMP signaling before and after mid-gastrulation [164]. Many transgenic lines have been created that show Smad-mediated BMP signaling in embryos and adults to visualize BMP signaling in live embryos: Tg(hsp70l:dnBmpr-GFP)w30, Tg(BmpRE-AAVmlp:eGFP)mw29, Tg(BRE-AAVmlp:d2GFP)mw30, Tg(BRE-AAVmlp:dmKO2)mw40, Tg(BMPRE:mRFP)cj100, and Tg(BMPRE:NLS-mCherry)ia17. These lines express eGFP, destabilized eGFP, destabilized KO2, or nuclear mCherry under the well-characterized BMP-responsive element (BRE), adopted from the mouse inhibitor of the differentiation-1 enhancer. The developing tailbud, hematopoietic lineage, dorsal eye, brain structures, heart, jaw, fins, and somite muscle, as well as other tissues, were found to dynamically express these fluorescent proteins [165,166,167,168].
To visualize CREB Signaling, Giuliodori et al. [169] generated a reporter line with a multimerized CREB-responsive element (6XCRE: 6X cAMP-responsive elements, 5′- TGACGTCA -3′). The transgenic reporter line is responsive through the transcription-factor-mediated activation of the regulatory elements located upstream of the coding sequences of the reporter protein [143]. Specifically, this reporter line shows CREB signaling activation in the heart and skeletal muscle, as well as in the nervous, hematopoietic, and endocrine systems (N. Tiso, personal communications).
Glucocorticoids (GCs) regulate many cellular processes including skeletal muscle development through the binding of the glucocorticoid receptor (GR) to specific REs located upstream of the transcription starting site or within an intron of GC target genes. The Dalla Valle group generated the transgenic line Tg(9xGCRE-HSV.Ul23:eGFP)ia20, in which GFP expression is driven by 9X GRE tandem repeats. The GRE is shared by the activated homodimerized receptors of GCs, mineralocorticoids, progesterone, and androgens [170]. This line shows a significant increase in eGFP fluorescence, starting with a widespread pattern during early somitogenesis and becoming primarily concentrated in the brain and trunk muscles 24 h post-fertilization [171].
Recently, the Moro group generated a novel Nrf2/ARE pathway biosensor fish, Tg(8xAORE:eGFP)ia201, which exhibits a dynamic spatiotemporal expression profile during early developmental stages in skeletal muscle. Nrf2 is a basic leucine zipper transcription factor that binds to the promoter region of the antioxidant response element (ARE), inducing the coordinated upregulation of antioxidant and detoxification genes. The Nrf2 reporter system, harboring eight tandem-repeated ARE sequences (5′-GTGACAAAGCA-3′), is derived from mouse and rat Glutathione-S-Transferase Alpha (Gsta) promoters [172]. High levels of fluorescence were detected in the gut, in muscle fibers, and in the caudal region. Since 7 dpf (or at the early larval stage), fluorescence has been also detected in motorneurons and in the heart [173].
Table 6. Zebrafish reporter lines targeting in vivo signaling pathway activities involved in muscle development.
Table 6. Zebrafish reporter lines targeting in vivo signaling pathway activities involved in muscle development.
LineTransgenic IDResponsive Element (RE)Signaling PathwayFeaturesRef.
Tg(hsp70l:dnBmpr-GFP)w30ZDB-ALT-050503-2BREBMP signalingHeat shock (hsp70) promoter GFP expression[164]
Tg(BmpRE:mRFP)cj100ZDB-ALT-110705-4BREBMP signalingMembrane-bound red fluorescent protein (RFP)[167,168]
Tg(BmpRE-AAVmlp:eGFP)mw29ZDB-ALT-110308-1BREBMP signalingGFP expression[166]
Tg(BRE-AAVmlp:d2GFP)mw30ZDB-ALT-110310-1BREBMP signalingDestabilized d2GFP[166]
Tg(BRE-AAVmlp:dmKO2)mw40ZDB-ALT-110310-2BREBMP signalingDestabilized monomeric Kusabira-Orange (dmKO2)[166]
Tg(BMPRE:NLS-mCherry)ia17ZDB-ALT-130115-2BREBMP signalingNuclear localization signal mCherry[165]
Tg(7xTCF-Xla.Siam:GFP)ia4ZDB-ALT-110113-1TCFWnt signalingGFP expression[163]
Tg(7xTCF-Xla.Siam:nlsmCherry)ia5ZDB-ALT-110113-2TCFWnt signalingNuclear localization signal mCherry[163]
Tg(TOP:GFP)w25ZDB-ALT-020621-4LEF binding sitesWnt signalingGFP expression[161,162]
Tg(6xCRE:eGFP) CRECREB signalingEnhanced GFP expression[169]
Tg(4xHRE-TATA:eGFP)ia21ZDB-ALT-131030-1HREHypoxia signalingEnhanced GFP expression[156]
Tg(12xgli-HSV.Ul23:GFP)ia11ZDB-ALT-120404-2gliShh signalingGFP expression[159]
Tg(8xGliBS:mCherry-NLS-Odc1)st1002ZDB-ALT-141030-2gliShh signalingNuclear-localized fluorescent protein tagged with an ornithine decarboxylase-derived destabilizing peptide[160]
Tg(9xGCRE-HSV.Ul23:eGFP)ia20 ZDB-ALT-130123-1 GREGlucocorticoidEnhanced GFP expression[171]
Tg(8XAORE:GFP)ia201ZDB-ALT-211202-1ARENrf2/AREGFP expression[173]

4.2. Zebrafish Reporter Lines Targeting the Skeletal Muscle

The zebrafish is a perfect vertebrate animal model to study the skeletal muscle in vivo both during development and disease progression. Furthermore, it is possible to create transgenic fish that may be used to label particular cell populations and monitor them over time, providing researchers with a rare chance to observe their live dynamics during development, damage, and regeneration. The different transgenic lines available for the live imaging of zebrafish muscle and stem/progenitor cells using confocal microscopes are provided in Table 7. Satellite cell markers that are well established and essential for muscle growth and regeneration include various transcription factors, including the paired homeobox genes pax3 and pax7, the myogenic regulatory factor (MRF) myf gene, the myogenin gene (myog), and myogenic differentiation 1 (myod) [174].
Serger et al. generated the transgenic lines TgBAC(pax7a:GFP)i131 and TgBAC(pax3a:GFP)i150 expressing fluorescent proteins under the control of the pax7a or pax3a promoters, and they are able to mark progenitors and satellite cells that contribute to the embryonic development and postembryonic growth of skeletal muscle, as well as its repair and regeneration [175]. Transgenic lines for myf5, myog, and myod have also been created to track muscle development in real-time [174,175,176].
The specification of muscle cell types in the zebrafish embryo is greatly influenced by Shh signaling activity [177]. Maurya et al. [178] describe the identification of a minimal element of the eng2a gene sufficient to drive reporter gene expression in the muscle progenitors and medial fast fiber cells.
The transgenic line Tg(-10en2a:eGFP)i233 integrates repressive and activating signals from the BMP and Shh pathways, respectively, to restrict the expression of eGFP in muscle progenitors and medial fast fibers.
The origin and mechanism of stem cell deployment or the clonal nature of muscle regeneration after injury could be more broadly assayed taking advantage of a muscle-specific lineage-tracing strategy using the zebrabow system, called “musclebow” [179]. The zebrabow line consists of a ubiquitous promoter driving three fluorophores (dTomato, mCerulean, and eYFP) in tandem and flanked by loxP sites. The construct’s default color is RFP (dTomato), but when Cre recombinase is present, the construct is stochastically deleted, allowing it to express either CFP (mCerulean) or YFP (eYFP). Since this excision is irreversible, the recombination event will be expressed in the cells’ offspring. As a result, cells and their offspring may be monitored over time. The achievement of muscle-specific recombination could be reached by crossing a muscle-specific CreERT2 line Tg(msgn1: CreERT2)pc9 with the Tg(ubi:Zebrabow)a131 line. Tamoxifen causes zebrabow rearrangements in all cells derived from the somites in these fish during embryogenesis because the mesogenin (msgn1) promoter is somite-specific [180]. These combinations of transgenes produce a “musclebow” fish, allowing the clonal relationships of muscle cells to be determined and resulting in individual fibers expressing distinctive color combinations. Newly differentiated fibers displayed a drift toward a single color (clonality) from the late larval stages until adulthood so that, in adults, entire sectors of each myotome have a uniform color. These findings show that multiple independent stem cells initially contribute randomly to fiber generation before a single stem cell clone eventually takes control of growth within individual myotomes over time due to repeated self-renewal events.
The transgenic line Tg(mylz2:GFP)gz8 with the mylz2 promoter is a specific molecular marker that monitors fast skeletal muscle development [181,182,183]. The GFP expression of the 1934-bp promoter transgenic lines closely mirrors the native mylz2 mRNA expression pattern in both somitic and non-somitic muscles, including the fin, eye, jaw, and gill muscles. Moreover, the Tg(tnnc1b:eGFP)i305 and Tg(smyhc1:GFP)i104 transgenic lines faithfully recapitulate the development of slow skeletal muscle.
Two transgenic lines, Tg(acta1:lifeact-GFP)pc21 and Tg(acta1:mCherryCAAX)pc22, are described by Berger et al. [184] to mark the myofibrils, sarcolemma, and t-tubules of the myofibers. In contrast, mCherry-CAAX integrates into the myofiber sarcolemma to fluorescently label the sarcolemma and t-tubules, while transgenic Lifeact-GFP highlights the striation of myofibrils. These two marker lines can be used together to examine and quantify the thickness of the myofibrils in individual live zebrafish myofibers [184,185]. In skeletal muscle disease, autophagy is a crucial cellular degradation system that transports cytoplasmic cargo to the lysosome [186]. Real-time autophagy monitoring is a crucial tool for understanding the formation of autophagosomes and determining the removal of proteins that are prone to aggregation. Two transgenic fluorescent zebrafish lines are used to monitor autophagy, one of which expresses RFP fused to LC3 Tg(hsp70l:lamp1-RFP)pd1064 and the other of which expresses RFP fused to lamp1b Tg(hsp70l:RFP-Rno.Map1lc3b)pd1065 under the control of the constitutive heat shock promoter [186,187].
Clark et al. created transgenic fluorescent zebrafish lines to create real-time and in vivo images of endosomes [188]. Based on the locations and roles of these endosome subtypes in zebrafish, scientists used Rab7 (late) to label late endosomes. These lines have been used to track autophagic activity in vivo and in real-time during development [186,187,188]. As previously described, zebrafish have many mitochondria in their skeletal muscles, which need high energy to contract and favor fish swimming. The balance of mitochondrial dynamics between fusion and fission keeps the web-shaped network in cells intact [189]. This balance is upset under specific physiological and pathological conditions, which leads to a change in mitochondrial morphology. The cytochrome c oxidase subunit VIIIA (cox8a) of zebrafish was cloned upstream of the eGFP gene by the Choi group [190]. Its expression was regulated by the widely used Xenopus EF-1 promoter. This line allows for visualizing mitochondria in motor neurons, erythroid cells, photoreceptor cells, and skeletal muscles to study mitochondrial morphology. Breeding these zebrafish transgenic lines to disease model zebrafish and subsequently observing their mitochondrial morphology as the disease develops would be of help in exploring, in vivo, how changes in mitochondrial morphology can activate caspase-mediated apoptosis affecting the pathogenesis of muscle diseases [190,191].
Table 7. Zebrafish reporter lines targeting the skeletal muscle.
Table 7. Zebrafish reporter lines targeting the skeletal muscle.
LineTransgenic IDPromoterCells TaggedFeaturesRef.
Tg(myf5:YFP)CLGY237
Tg(−80.0myf5: EGFP)zf37
ZDB-ALT-150512-2 ZDB-ALT-070730-1Myogenic factor 5Quiescent and activated satellite cellsYellow fluorescent protein
Enhanced GFP expression
[176,192]
TgBAC(pax7a:GFP)i131ZDB-ALT-111118-32Paired box 7aMuscle progenitorsGFP expression[175]
TgBAC(pax3a:GFP)i150ZDB-ALT-111118-34Paired box 3aMuscle progenitors GFP expression[175]
Tg(myog:GFP)pc27ZDB-ALT-171003-14MyogeninMuscle progenitorseGFP expression[174]
TgBAC(myod:GFP)i124ZDB-ALT-111118-35Myogenic differentiation 1Muscle progenitorsGFP expression[175]
Tg(-10en2a:EGFP)i233ZDB-ALT-110223-3Engrailed homeobox 2aMuscle progenitors and medial fast fibersEnhanced GFP expression[178]
Tg(-2.2mylz2:GFP)i135
Tg(mylz2:GFP)gz8
Tg(mylpfa:mCherry)cz3327
ZDB-ALT-081112-1
ZDB-ALT-080207-2
ZDB-ALT-130923-2
Myosin light chain, phosphorylatable, fast skeletal muscleFast skeletal muscleGFP expression
or mCherry
[181,182,183]
TgPAC(prdm1a:EGFP)i106 ZDB-ALT-080923-6 PR domain containing 1a, with ZNF domain Primary and secondary slow fibersEnhanced GFP expression[193]
Tg(tnnc1b:eGFP)i305ZDB-ALT-150723-4Troponin C type 1bSlow-twitch muscle fibersEnhanced GFP expression[146]
Tg(Tru.Myhz1.1:EGFP)kj100ZDB-ALT-130823-1 Myosin, heavy polypeptide 1.1, skeletal muscle Slow muscle-specific heavy chainEnhanced GFP expression[194]
Tg(smyhc1:GFP)i104
Tg(smyhc1:LY-Tomato)oz29
ZDB-ALT-080923-4
ZDB-ALT-191024-1
Slow myosin heavy chain 1 Slow-twitch muscle fibers GFP or tomato expression[146,193,195]
Tg(acta1: mCherryCAAX)pc22
TgBAC(actc1b: GFP)zf13: zf13Tg
Tg(acta1:lifeact-GFP)pc21
ZDB-ALT-150224-2
ZDB-ALT-060221-2
ZDB-ALT-150224-1
Actin alpha cardiac muscle 1bMyofibrils, sarcolemma, and t-tubules of the myofibersMembrane-tethered mCherry or GFP[184,185]
Tg(ubi: zebrabow-M)a131
Tg(msgn1: CreERT2)pc9
ZDB-ALT-130816-2
ZDB-ALT-141117-7
Mesogenin 1Muscle progenitors dTomato, mCerulean and eYFP) [196,197]
Tg(hsp70l:lamp1-RFP)pd1064 ZDB-ALT-130409-7 Lysosomal-associated membrane protein 1b Skeletal muscle
lysosomes
RFP expression[186,187]
Tg(hsp70l:RFP-Rno.Map1lc3b)pd1065 ZDB-ALT-130410-5 Microtubule-associated protein 1 light chain 3 beta Skeletal muscle
lysosomes
RFP expression[186,187]
Tg(h2ax:EGFP-rab7a)mw7 ZDB-ALT-111017-6 RAB7a, member RAS oncogene family Skeletal muscle
Late endosome
Enhanced GFP expression[188]
Tg(Xla.Eef1a1:mlsEGFP) cms1 ZDB-ALT-090309-2 Cytochrome c oxidase subunit 8A MitochondriaEnhanced GFP expression[190,191]

5. Zebrafish Models of MDs Used for Drug Screening

Nonmammalian animal models, such as Caenorabditis elegans, Drosophyla melanogaster, and Danio rerio, are particularly attractive in conducting cost-effective, high-throughput drug screenings. This is because of the ease of genetic manipulation, high offspring numbers, and low ethical concerns compared with existing mammalian models such as mdx mice and Golden Retriever dogs [198]. The major issue of mdx mice, which is the most used animal model for DMD, is the mild clinical presentation due to compensatory phenomena [199,200]. For that reason, several mdx mouse derivatives have been developed to have a disease phenotype more similar to humans and Golden Retrievers [201,202]. Among nonmammalian models, zebrafish models of myopathies show severe disease phenotypes and have been extensively used for high-throughput drug screening. A second advantage of zebrafish models is that embryos can easily absorb drugs dissolved in embryo water, which is useful in assessing the effects during the first stages of disease development [203]. Several lead compounds able to ameliorate major disease symptoms, such as muscle fibrosis, inflammation, dysregulation of calcium homeostasis, oxidative stress, and mitochondrial dysfunction, have been discovered by using DMD sapje and sapje-like zebrafish models generated with ENU mutagenesis (see Table 2) [117,118,137,198,204,205,206,207,208], LGMD2I zebrafish models harboring the fkrp mutation (see Table 3) [76], the zebrafish transgenic model of myotonic dystrophy harboring multiple CUG repetitions in its genome [209], zebrafish models of collagen VI deficiency congenital myopathies (Table 1) [119,210,211], and candyfloss zebrafish models [33].

6. Conclusions

Zebrafish mutant lines, transgenic fluorescent biosensors, and phenotyping assays that have been extensively used in biomedical research to study muscle development and perform high-throughput drug screening were summarized in this review. All these tools are essential to understanding the molecular mechanisms underlying both the pathogenesis of muscular dystrophies and skeletal muscle development and organization. Muscle-specific transgenic zebrafish biosensors are a powerful tool to track the cell dynamics in a living animal at the level of a single muscle fiber at the embryonic or larval stages. Thanks to the advances in new technologies, such as CRISPR/Cas9, we will be able to both generate even more specific zebrafish lines mimicking human muscular dystrophies and to progress toward precision medicine by repairing muscle defects in patients showing specific mutations in specific muscle genes.

Author Contributions

Conceptualization, N.F., M.S. and A.V.; software, C.T.; validation, C.T., N.F., M.S. and A.V.; formal analysis, N.F., M.S. and A.V.; investigation, C.T., F.G. (Francesca Greco), E.C., F.G. (Francesco Ghirotto), N.F., M.S. and A.V.; resources, N.F., M.S. and A.V.; data curation, N.F., M.S. and A.V.; writing—original draft preparation, C.T., F.G. (Francesca Greco), E.C., F.G. (Francesco Ghirotto), N.F., M.S. and A.V.; visualization, C.T., F.G. (Francesca Greco), E.C., F.G. (Francesco Ghirotto), N.F., M.S. and A.V.; supervision, N.F., M.S. and A.V.; project administration, N.F., M.S. and A.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Non applicable.

Informed Consent Statement

Non applicable.

Data Availability Statement

Non applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Davies, K.E.; Nowak, K.J. Molecular mechanisms of muscular dystrophies: Old and new players. Nat. Rev. Mol. Cell Biol. 2006, 7, 762–773. [Google Scholar] [CrossRef] [PubMed]
  2. Carter, J.C.; Sheehan, D.W.; Prochoroff, A.; Birnkrant, D.J. Muscular Dystrophies. Clin. Chest Med. 2018, 39, 377–389. [Google Scholar] [CrossRef] [PubMed]
  3. Eriksson, M.; Brown, W.T.; Gordon, L.B.; Glynn, M.W.; Singer, J.; Scott, L.; Erdos, M.R.; Robbins, C.M.; Moses, T.Y.; Berglund, P.; et al. Recurrent de novo point mutations in lamin A cause Hutchinson-Gilford progeria syndrome. Nature 2003, 423, 293–298. [Google Scholar] [CrossRef]
  4. Hussain, I.; Patni, N.; Ueda, M.; Sorkina, E.; Valerio, C.M.; Cochran, E.; Brown, R.J.; Peeden, J.; Tikhonovich, Y.; Tiulpakov, A.; et al. A Novel Generalized Lipodystrophy-Associated Progeroid Syndrome Due to Recurrent Heterozygous LMNA p.T10I Mutation. J. Clin. Endocrinol. Metab. 2018, 103, 1005–1014. [Google Scholar] [CrossRef] [PubMed]
  5. Yukina, M.; Nuralieva, N.; Sorkina, E.; Troshina, E.; Tiulpakov, A.; Belaya, Z.; Melnichenko, G. Atypical progeroid syndrome (p.E262K LMNA mutation): A rare cause of short stature and osteoporosis. Endocrinol. Diabetes Metab. Case Rep. 2021, 2021, 20-0188. [Google Scholar] [CrossRef]
  6. Lieschke, G.J.; Currie, P.D. Animal models of human disease: Zebrafish swim into view. Nat. Rev. Genet. 2007, 8, 353–367. [Google Scholar] [CrossRef]
  7. 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. [Google Scholar] [CrossRef]
  8. Hall, T.E.; Bryson-Richardson, R.J.; Berger, S.; Jacoby, A.S.; Cole, N.J.; Hollway, G.E.; Berger, J.; Currie, P.D. The zebrafish candyfloss mutant implicates extracellular matrix adhesion failure in laminin alpha2-deficient congenital muscular dystrophy. Proc. Natl. Acad. Sci. USA 2007, 104, 7092–7097. [Google Scholar] [CrossRef]
  9. Vacaru, A.M.; Unlu, G.; Spitzner, M.; Mione, M.; Knapik, E.W.; Sadler, K.C. In vivo cell biology in zebrafish—Providing insights into vertebrate development and disease. J. Cell Sci. 2014, 127 Pt 3, 485–495. [Google Scholar] [CrossRef]
  10. Frontera, W.R.; Ochala, J. Skeletal muscle: A brief review of structure and function. Calcif. Tissue Int. 2015, 96, 183–195. [Google Scholar] [CrossRef]
  11. Charvet, B.; Malbouyres, M.; Pagnon-Minot, A.; Ruggiero, F.; Le Guellec, D. Development of the zebrafish myoseptum with emphasis on the myotendinous junction. Cell Tissue Res. 2011, 346, 439–449. [Google Scholar] [CrossRef] [PubMed]
  12. Henry, C.A.; McNulty, I.M.; Durst, W.A.; Munchel, S.E.; Amacher, S.L. Interactions between muscle fibers and segment boundaries in zebrafish. Dev. Biol. 2005, 287, 346–360. [Google Scholar] [CrossRef] [PubMed]
  13. Stickney, H.L.; Barresi, M.J.; Devoto, S.H. Somite development in zebrafish. Dev. Dyn. 2000, 219, 287–303. [Google Scholar] [CrossRef] [PubMed]
  14. Hollway, G.E.; Bryson-Richardson, R.J.; Berger, S.; Cole, N.J.; Hall, T.E.; Currie, P.D. Whole-somite rotation generates muscle progenitor cell compartments in the developing zebrafish embryo. Dev. Cell 2007, 12, 207–219. [Google Scholar] [CrossRef] [PubMed]
  15. Altringham, J.D.; Ellerby, D.J. Fish swimming: Patterns in muscle function. J. Exp. Biol. 1999, 202 Pt 23, 3397–3403. [Google Scholar] [CrossRef] [PubMed]
  16. Squire, J.M. Architecture and function in the muscle sarcomere. Curr. Opin. Struct. Biol. 1997, 7, 247–257. [Google Scholar] [CrossRef]
  17. Dube, D.K.; Dube, S.; Abbott, L.; Wang, J.; Fan, Y.; Alshiekh-Nasany, R.; Shah, K.K.; Rudloff, A.P.; Poiesz, B.J.; Sanger, J.M.; et al. Identification, characterization, and expression of sarcomeric tropomyosin isoforms in zebrafish. Cytoskeleton 2017, 74, 125–142. [Google Scholar] [CrossRef]
  18. Ervasti, J.M.; Campbell, K.P. A role for the dystrophin-glycoprotein complex as a transmembrane linker between laminin and actin. J. Cell Biol. 1993, 122, 809–823. [Google Scholar] [CrossRef]
  19. Lapidos, K.A.; Kakkar, R.; McNally, E.M. The dystrophin glycoprotein complex: Signaling strength and integrity for the sarcolemma. Circ. Res. 2004, 94, 1023–1031. [Google Scholar] [CrossRef] [PubMed]
  20. Barresi, R.; Campbell, K.P. Dystroglycan: From biosynthesis to pathogenesis of human disease. J. Cell Sci. 2006, 119 Pt 2, 199–207. [Google Scholar] [CrossRef] [PubMed]
  21. Emery, A.E. The muscular dystrophies. Lancet 2002, 359, 687–695. [Google Scholar] [CrossRef]
  22. Muntoni, F.; Voit, T. The congenital muscular dystrophies in 2004: A century of exciting progress. Neuromuscul. Disord. 2004, 14, 635–649. [Google Scholar] [CrossRef] [PubMed]
  23. Nigro, V.; de Sa Moreira, E.; Piluso, G.; Vainzof, M.; Belsito, A.; Politano, L.; Puca, A.A.; Passos-Bueno, M.R.; Zatz, M. Autosomal recessive limb-girdle muscular dystrophy, LGMD2F, is caused by a mutation in the delta-sarcoglycan gene. Nat. Genet. 1996, 14, 195–198. [Google Scholar] [CrossRef] [PubMed]
  24. Helbling-Leclerc, A.; Zhang, X.; Topaloglu, H.; Cruaud, C.; Tesson, F.; Weissenbach, J.; Tome, F.M.; Schwartz, K.; Fardeau, M.; Tryggvason, K.; et al. Mutations in the laminin alpha 2-chain gene (LAMA2) cause merosin-deficient congenital muscular dystrophy. Nat. Genet. 1995, 11, 216–218. [Google Scholar] [CrossRef] [PubMed]
  25. Moreira, E.S.; Wiltshire, T.J.; Faulkner, G.; Nilforoushan, A.; Vainzof, M.; Suzuki, O.T.; Valle, G.; Reeves, R.; Zatz, M.; Passos-Bueno, M.R.; et al. Limb-girdle muscular dystrophy type 2G is caused by mutations in the gene encoding the sarcomeric protein telethonin. Nat. Genet. 2000, 24, 163–166. [Google Scholar] [CrossRef]
  26. Anderson, L.V.; Davison, K.; Moss, J.A.; Young, C.; Cullen, M.J.; Walsh, J.; Johnson, M.A.; Bashir, R.; Britton, S.; Keers, S.; et al. Dysferlin is a plasma membrane protein and is expressed early in human development. Hum. Mol. Genet. 1999, 8, 855–861. [Google Scholar] [CrossRef]
  27. Bertini, E.; Pepe, G. Collagen type VI and related disorders: Bethlem myopathy and Ullrich scleroatonic muscular dystrophy. Eur. J. Paediatr. Neurol. 2002, 6, 193–198. [Google Scholar] [CrossRef]
  28. Richard, I.; Broux, O.; Allamand, V.; Fougerousse, F.; Chiannilkulchai, N.; Bourg, N.; Brenguier, L.; Devaud, C.; Pasturaud, P.; Roudaut, C.; et al. Mutations in the proteolytic enzyme calpain 3 cause limb-girdle muscular dystrophy type 2A. Cell 1995, 81, 27–40. [Google Scholar] [CrossRef]
  29. Genschel, J.; Schmidt, H.H. Mutations in the LMNA gene encoding lamin A/C. Hum. Mutat. 2000, 16, 451–459. [Google Scholar] [CrossRef]
  30. Mahdy, M.A.A. Skeletal muscle fibrosis: An overview. Cell Tissue Res. 2019, 375, 575–588. [Google Scholar] [CrossRef]
  31. Van Ry, P.M.; Fontelonga, T.M.; Barraza-Flores, P.; Sarathy, A.; Nunes, A.M.; Burkin, D.J. ECM-Related Myopathies and Muscular Dystrophies: Pros and Cons of Protein Therapies. Compr. Physiol. 2017, 7, 1519–1536. [Google Scholar] [PubMed]
  32. Zambon, A.A.; Muntoni, F. Congenital muscular dystrophies: What is new? Neuromuscul. Disord. 2021, 31, 931–942. [Google Scholar] [CrossRef] [PubMed]
  33. Smith, S.J.; Wang, J.C.; Gupta, V.A.; Dowling, J.J. A novel early onset phenotype in a zebrafish model of merosin deficient congenital muscular dystrophy. PLoS ONE 2017, 12, e0172648. [Google Scholar] [CrossRef] [PubMed]
  34. Li, M.; Arner, A. Immobilization of Dystrophin and Laminin alpha2-Chain Deficient Zebrafish Larvae In Vivo Prevents the Development of Muscular Dystrophy. PLoS ONE 2015, 10, e0139483. [Google Scholar]
  35. Hall, T.E.; Wood, A.J.; Ehrlich, O.; Li, M.; Sonntag, C.S.; Cole, N.J.; Huttner, I.G.; Sztal, T.E.; Currie, P.D. Cellular rescue in a zebrafish model of congenital muscular dystrophy type 1A. npj Regen. Med. 2019, 4, 21. [Google Scholar] [CrossRef] [PubMed]
  36. Smith, S.J.; Horstick, E.J.; Davidson, A.E.; Dowling, J. Analysis of Zebrafish Larvae Skeletal Muscle Integrity with Evans Blue Dye. J. Vis. Exp. 2015, e53183. [Google Scholar] [CrossRef]
  37. Smith, S.J.; Fabian, L.; Sheikh, A.; Noche, R.; Cui, X.; Moore, S.A.; Dowling, J.J. Lysosomes and the pathogenesis of merosin-deficient congenital muscular dystrophy. Hum. Mol. Genet. 2022, 31, 733–747. [Google Scholar] [CrossRef]
  38. Gupta, V.A.; Kawahara, G.; Myers, J.A.; Chen, A.T.; Hall, T.E.; Manzini, M.C.; Currie, P.D.; Zhou, Y.; Zon, L.I.; Kunkel, L.M.; et al. A splice site mutation in laminin-alpha2 results in a severe muscular dystrophy and growth abnormalities in zebrafish. PLoS ONE 2012, 7, e43794. [Google Scholar] [CrossRef]
  39. Sztal, T.E.; Sonntag, C.; Hall, T.E.; Currie, P.D. Epistatic dissection of laminin-receptor interactions in dystrophic zebrafish muscle. Hum. Mol. Genet. 2012, 21, 4718–4731. [Google Scholar] [CrossRef]
  40. Fabian, L.; Dowling, J.J. Zebrafish Models of LAMA2-Related Congenital Muscular Dystrophy (MDC1A). Front. Mol. Neurosci. 2020, 13, 122. [Google Scholar] [CrossRef]
  41. Mienaltowski, M.J.; Birk, D.E. Structure, physiology, and biochemistry of collagens. Adv. Exp. Med. Biol. 2014, 802, 5–29. [Google Scholar]
  42. Lampe, A.K.; Bushby, K.M. Collagen VI related muscle disorders. J. Med. Genet. 2005, 42, 673–685. [Google Scholar] [CrossRef] [PubMed]
  43. Charvet, B.; Guiraud, A.; Malbouyres, M.; Zwolanek, D.; Guillon, E.; Bretaud, S.; Monnot, C.; Schulze, J.; Bader, H.L.; Allard, B.; et al. Knockdown of col22a1 gene in zebrafish induces a muscular dystrophy by disruption of the myotendinous junction. Development 2013, 140, 4602–4613. [Google Scholar] [CrossRef] [PubMed]
  44. Tonelotto, V.; Consorti, C.; Facchinello, N.; Trapani, V.; Sabatelli, P.; Giraudo, C.; Spizzotin, M.; Cescon, M.; Bertolucci, C.; Bonaldo, P. Collagen VI ablation in zebrafish causes neuromuscular defects during developmental and adult stages. Matrix Biol. 2022, 112, 39–61. [Google Scholar] [CrossRef] [PubMed]
  45. Radev, Z.; Hermel, J.M.; Elipot, Y.; Bretaud, S.; Arnould, S.; Duchateau, P.; Ruggiero, F.; Joly, J.S.; Sohm, F. A TALEN-Exon Skipping Design for a Bethlem Myopathy Model in Zebrafish. PLoS ONE 2015, 10, e0133986. [Google Scholar] [CrossRef]
  46. Telfer, W.R.; Busta, A.S.; Bonnemann, C.G.; Feldman, E.L.; Dowling, J.J. Zebrafish models of collagen VI-related myopathies. Hum. Mol. Genet. 2010, 19, 2433–2444. [Google Scholar] [CrossRef]
  47. Malbouyres, M.; Guiraud, A.; Lefrancois, C.; Salamito, M.; Nauroy, P.; Bernard, L.; Sohm, F.; Allard, B.; Ruggiero, F. Lack of the myotendinous junction marker col22a1 results in posture and locomotion disabilities in zebrafish. Matrix Biol. 2022, 109, 1–18. [Google Scholar] [CrossRef]
  48. Parsons, M.J.; Campos, I.; Hirst, E.M.; Stemple, D.L. Removal of dystroglycan causes severe muscular dystrophy in zebrafish embryos. Development 2002, 129, 3505–3512. [Google Scholar] [CrossRef]
  49. Bassett, D.I.; Bryson-Richardson, R.J.; Daggett, D.F.; Gautier, P.; Keenan, D.G.; Currie, P.D. Dystrophin is required for the formation of stable muscle attachments in the zebrafish embryo. Development 2003, 130, 5851–5860. [Google Scholar] [CrossRef]
  50. Granato, M.; van Eeden, F.J.; Schach, U.; Trowe, T.; Brand, M.; Furutani-Seiki, M.; Haffter, P.; Hammerschmidt, M.; Heisenberg, C.P.; Jiang, Y.J.; et al. Genes controlling and mediating locomotion behavior of the zebrafish embryo and larva. Development 1996, 123, 399–413. [Google Scholar] [CrossRef]
  51. Bolanos-Jimenez, F.; Bordais, A.; Behra, M.; Strahle, U.; Sahel, J.; Rendon, A. Dystrophin and Dp71, two products of the DMD gene, show a different pattern of expression during embryonic development in zebrafish. Mech. Dev. 2001, 102, 239–241. [Google Scholar] [CrossRef] [PubMed]
  52. Guyon, J.R.; Goswami, J.; Jun, S.J.; Thorne, M.; Howell, M.; Pusack, T.; Kawahara, G.; Steffen, L.S.; Galdzicki, M.; Kunkel, L.M. Genetic isolation and characterization of a splicing mutant of zebrafish dystrophin. Hum. Mol. Genet. 2009, 18, 202–211. [Google Scholar] [CrossRef] [PubMed]
  53. Mitsuhashi, H.; Mitsuhashi, S.; Lynn-Jones, T.; Kawahara, G.; Kunkel, L.M. Expression of DUX4 in zebrafish development recapitulates facioscapulohumeral muscular dystrophy. Hum. Mol. Genet. 2013, 22, 568–577. [Google Scholar] [CrossRef]
  54. Pakula, A.; Lek, A.; Widrick, J.; Mitsuhashi, H.; Bugda Gwilt, K.M.; Gupta, V.A.; Rahimov, F.; Criscione, J.; Zhang, Y.; Gibbs, D.; et al. Transgenic zebrafish model of DUX4 misexpression reveals a developmental role in FSHD pathogenesis. Hum. Mol. Genet. 2019, 28, 320–331. [Google Scholar] [CrossRef] [PubMed]
  55. Wallace, L.M.; Garwick, S.E.; Mei, W.; Belayew, A.; Coppee, F.; Ladner, K.J.; Guttridge, D.; Yang, J.; Harper, S.Q. DUX4, a candidate gene for facioscapulohumeral muscular dystrophy, causes p53-dependent myopathy in vivo. Ann. Neurol. 2011, 69, 540–552. [Google Scholar] [CrossRef]
  56. Hinman, M.N.; Richardson, J.I.; Sockol, R.A.; Aronson, E.D.; Stednitz, S.J.; Murray, K.N.; Berglund, J.A.; Guillemin, K. Zebrafish mbnl mutants model physical and molecular phenotypes of myotonic dystrophy. Dis. Model Mech. 2021, 14, dmm045773. [Google Scholar] [CrossRef]
  57. Machuca-Tzili, L.E.; Buxton, S.; Thorpe, A.; Timson, C.M.; Wigmore, P.; Luther, P.K.; Brook, J.D. Zebrafish deficient for Muscleblind-like 2 exhibit features of myotonic dystrophy. Dis. Model Mech. 2011, 4, 381–392. [Google Scholar] [CrossRef]
  58. Todd, P.K.; Ackall, F.Y.; Hur, J.; Sharma, K.; Paulson, H.L.; Dowling, J.J. Transcriptional changes and developmental abnormalities in a zebrafish model of myotonic dystrophy type 1. Dis. Model Mech. 2014, 7, 143–155. [Google Scholar] [CrossRef]
  59. Kustermann, M.; Manta, L.; Paone, C.; Kustermann, J.; Lausser, L.; Wiesner, C.; Eichinger, L.; Clemen, C.S.; Schroder, R.; Kestler, H.A.; et al. Loss of the novel Vcp (valosin containing protein) interactor Washc4 interferes with autophagy-mediated proteostasis in striated muscle and leads to myopathy in vivo. Autophagy 2018, 14, 1911–1927. [Google Scholar] [CrossRef]
  60. Voisard, P.; Diofano, F.; Glazier, A.A.; Rottbauer, W.; Just, S. CRISPR/Cas9-Mediated Constitutive Loss of VCP (Valosin-Containing Protein) Impairs Proteostasis and Leads to Defective Striated Muscle Structure and Function In Vivo. Int. J. Mol. Sci. 2022, 23, 6722. [Google Scholar] [CrossRef]
  61. Buhrdel, J.B.; Hirth, S.; Kessler, M.; Westphal, S.; Forster, M.; Manta, L.; Wiche, G.; Schoser, B.; Schessl, J.; Schroder, R.; et al. In vivo characterization of human myofibrillar myopathy genes in zebrafish. Biochem. Biophys. Res. Commun. 2015, 461, 217–223. [Google Scholar] [CrossRef] [PubMed]
  62. Guyon, J.R.; Mosley, A.N.; Zhou, Y.; O’Brien, K.F.; Sheng, X.; Chiang, K.; Davidson, A.J.; Volinski, J.M.; Zon, L.I.; Kunkel, L.M. The dystrophin associated protein complex in zebrafish. Hum. Mol. Genet. 2003, 12, 601–615. [Google Scholar] [CrossRef] [PubMed]
  63. Lo, H.P.; Nixon, S.J.; Hall, T.E.; Cowling, B.S.; Ferguson, C.; Morgan, G.P.; Schieber, N.L.; Fernandez-Rojo, M.A.; Bastiani, M.; Floetenmeyer, M.; et al. The caveolin-cavin system plays a conserved and critical role in mechanoprotection of skeletal muscle. J. Cell Biol. 2015, 210, 833–849. [Google Scholar] [CrossRef] [PubMed]
  64. Vieira, N.M.; Naslavsky, M.S.; Licinio, L.; Kok, F.; Schlesinger, D.; Vainzof, M.; Sanchez, N.; Kitajima, J.P.; Gal, L.; Cavacana, N.; et al. A defect in the RNA-processing protein HNRPDL causes limb-girdle muscular dystrophy 1G (LGMD1G). Hum. Mol. Genet. 2014, 23, 4103–4110. [Google Scholar] [CrossRef]
  65. Zhang, R.; Yang, J.; Zhu, J.; Xu, X. Depletion of zebrafish Tcap leads to muscular dystrophy via disrupting sarcomere-membrane interaction, not sarcomere assembly. Hum. Mol. Genet. 2009, 18, 4130–4140. [Google Scholar] [CrossRef]
  66. Steffen, L.S.; Guyon, J.R.; Vogel, E.D.; Howell, M.H.; Zhou, Y.; Weber, G.J.; Zon, L.I.; Kunkel, L.M. The zebrafish runzel muscular dystrophy is linked to the titin gene. Dev. Biol. 2007, 309, 180–192. [Google Scholar] [CrossRef]
  67. Schindler, R.F.; Scotton, C.; Zhang, J.; Passarelli, C.; Ortiz-Bonnin, B.; Simrick, S.; Schwerte, T.; Poon, K.L.; Fang, M.; Rinne, S.; et al. POPDC1(S201F) causes muscular dystrophy and arrhythmia by affecting protein trafficking. J. Clin. Investig. 2016, 126, 239–253. [Google Scholar] [CrossRef]
  68. Osborn, D.P.S.; Pond, H.L.; Mazaheri, N.; Dejardin, J.; Munn, C.J.; Mushref, K.; Cauley, E.S.; Moroni, I.; Pasanisi, M.B.; Sellars, E.A.; et al. Mutations in INPP5K Cause a Form of Congenital Muscular Dystrophy Overlapping Marinesco-Sjogren Syndrome and Dystroglycanopathy. Am. J. Hum. Genet. 2017, 100, 537–545. [Google Scholar] [CrossRef]
  69. Wiessner, M.; Roos, A.; Munn, C.J.; Viswanathan, R.; Whyte, T.; Cox, D.; Schoser, B.; Sewry, C.; Roper, H.; Phadke, R.; et al. Mutations in INPP5K, Encoding a Phosphoinositide 5-Phosphatase, Cause Congenital Muscular Dystrophy with Cataracts and Mild Cognitive Impairment. Am. J. Hum. Genet. 2017, 100, 523–536. [Google Scholar] [CrossRef]
  70. Ruparelia, A.A.; Oorschot, V.; Vaz, R.; Ramm, G.; Bryson-Richardson, R.J. Zebrafish models of BAG3 myofibrillar myopathy suggest a toxic gain of function leading to BAG3 insufficiency. Acta Neuropathol. 2014, 128, 821–833. [Google Scholar] [CrossRef]
  71. Kessler, M.; Kieltsch, A.; Kayvanpour, E.; Katus, H.A.; Schoser, B.; Schessl, J.; Just, S.; Rottbauer, W. A zebrafish model for FHL1-opathy reveals loss-of-function effects of human FHL1 mutations. Neuromuscul. Disord. 2018, 28, 521–531. [Google Scholar] [CrossRef]
  72. van der Meer, D.L.; Marques, I.J.; Leito, J.T.; Besser, J.; Bakkers, J.; Schoonheere, E.; Bagowski, C.P. Zebrafish cypher is important for somite formation and heart development. Dev. Biol. 2006, 299, 356–372. [Google Scholar] [CrossRef]
  73. Jahncke, J.N.; Wright, K.M. The many roles of dystroglycan in nervous system development and function: Dystroglycan and neural circuit development: Dystroglycan and neural circuit development. Dev. Dyn. 2023, 252, 61–80. [Google Scholar] [CrossRef] [PubMed]
  74. Vannoy, C.H.; Leroy, V.; Lu, Q.L. Dose-Dependent Effects of FKRP Gene-Replacement Therapy on Functional Rescue and Longevity in Dystrophic Mice. Mol. Ther. Methods Clin. Dev. 2018, 11, 106–120. [Google Scholar] [CrossRef]
  75. Gupta, V.; Kawahara, G.; Gundry, S.R.; Chen, A.T.; Lencer, W.I.; Zhou, Y.; Zon, L.I.; Kunkel, L.M.; Beggs, A.H. The zebrafish dag1 mutant: A novel genetic model for dystroglycanopathies. Hum. Mol. Genet. 2011, 20, 1712–1725. [Google Scholar] [CrossRef]
  76. Serafini, P.R.; Feyder, M.J.; Hightower, R.M.; Garcia-Perez, D.; Vieira, N.M.; Lek, A.; Gibbs, D.E.; Moukha-Chafiq, O.; Augelli-Szafran, C.E.; Kawahara, G.; et al. A limb-girdle muscular dystrophy 2I model of muscular dystrophy identifies corrective drug compounds for dystroglycanopathies. JCI Insight 2018, 3, e120493. [Google Scholar] [CrossRef] [PubMed]
  77. Wood, A.J.; Lin, C.H.; Li, M.; Nishtala, K.; Alaei, S.; Rossello, F.; Sonntag, C.; Hersey, L.; Miles, L.B.; Krisp, C.; et al. FKRP-dependent glycosylation of fibronectin regulates muscle pathology in muscular dystrophy. Nat. Commun. 2021, 12, 2951. [Google Scholar] [CrossRef] [PubMed]
  78. Avsar-Ban, E.; Ishikawa, H.; Manya, H.; Watanabe, M.; Akiyama, S.; Miyake, H.; Endo, T.; Tamaru, Y. Protein O-mannosylation is necessary for normal embryonic development in zebrafish. Glycobiology 2010, 20, 1089–1102. [Google Scholar] [CrossRef]
  79. Manzini, M.C.; Tambunan, D.E.; Hill, R.S.; Yu, T.W.; Maynard, T.M.; Heinzen, E.L.; Shianna, K.V.; Stevens, C.R.; Partlow, J.N.; Barry, B.J.; et al. Exome sequencing and functional validation in zebrafish identify GTDC2 mutations as a cause of Walker-Warburg syndrome. Am. J. Hum. Genet. 2012, 91, 541–547. [Google Scholar] [CrossRef]
  80. Lin, Y.Y.; White, R.J.; Torelli, S.; Cirak, S.; Muntoni, F.; Stemple, D.L. Zebrafish Fukutin family proteins link the unfolded protein response with dystroglycanopathies. Hum. Mol. Genet. 2011, 20, 1763–1775. [Google Scholar] [CrossRef]
  81. Wood, A.J.; Muller, J.S.; Jepson, C.D.; Laval, S.H.; Lochmuller, H.; Bushby, K.; Barresi, R.; Straub, V. Abnormal vascular development in zebrafish models for fukutin and FKRP deficiency. Hum. Mol. Genet. 2011, 20, 4879–4890. [Google Scholar] [CrossRef] [PubMed]
  82. Thornhill, P.; Bassett, D.; Lochmuller, H.; Bushby, K.; Straub, V. Developmental defects in a zebrafish model for muscular dystrophies associated with the loss of fukutin-related protein (FKRP). Brain 2008, 131 Pt 6, 1551–1561. [Google Scholar] [CrossRef]
  83. Kawahara, G.; Guyon, J.R.; Nakamura, Y.; Kunkel, L.M. Zebrafish models for human FKRP muscular dystrophies. Hum. Mol. Genet. 2010, 19, 623–633. [Google Scholar] [CrossRef] [PubMed]
  84. Roscioli, T.; Kamsteeg, E.J.; Buysse, K.; Maystadt, I.; van Reeuwijk, J.; van den Elzen, C.; van Beusekom, E.; Riemersma, M.; Pfundt, R.; Vissers, L.E.; et al. Mutations in ISPD cause Walker-Warburg syndrome and defective glycosylation of alpha-dystroglycan. Nat. Genet. 2012, 44, 581–585. [Google Scholar] [CrossRef] [PubMed]
  85. Praissman, J.L.; Willer, T.; Sheikh, M.O.; Toi, A.; Chitayat, D.; Lin, Y.Y.; Lee, H.; Stalnaker, S.H.; Wang, S.; Prabhakar, P.K.; et al. The functional O-mannose glycan on alpha-dystroglycan contains a phospho-ribitol primed for matriglycan addition. eLife 2016, 5, e14473. [Google Scholar] [CrossRef]
  86. Stevens, E.; Carss, K.J.; Cirak, S.; Foley, A.R.; Torelli, S.; Willer, T.; Tambunan, D.E.; Yau, S.; Brodd, L.; Sewry, C.A.; et al. Mutations in B3GALNT2 cause congenital muscular dystrophy and hypoglycosylation of alpha-dystroglycan. Am. J. Hum. Genet. 2013, 92, 354–365. [Google Scholar] [CrossRef]
  87. Buysse, K.; Riemersma, M.; Powell, G.; van Reeuwijk, J.; Chitayat, D.; Roscioli, T.; Kamsteeg, E.J.; van den Elzen, C.; van Beusekom, E.; Blaser, S.; et al. Missense mutations in beta-1,3-N-acetylglucosaminyltransferase 1 (B3GNT1) cause Walker-Warburg syndrome. Hum. Mol. Genet. 2013, 22, 1746–1754. [Google Scholar] [CrossRef]
  88. Ardiccioni, C.; Clarke, O.B.; Tomasek, D.; Issa, H.A.; von Alpen, D.C.; Pond, H.L.; Banerjee, S.; Rajashankar, K.R.; Liu, Q.; Guan, Z.; et al. Structure of the polyisoprenyl-phosphate glycosyltransferase GtrB and insights into the mechanism of catalysis. Nat. Commun. 2016, 7, 10175. [Google Scholar] [CrossRef]
  89. Marchese, M.; Pappalardo, A.; Baldacci, J.; Verri, T.; Doccini, S.; Cassandrini, D.; Bruno, C.; Fiorillo, C.; Garcia-Gil, M.; Bertini, E.; et al. Dolichol-phosphate mannose synthase depletion in zebrafish leads to dystrophic muscle with hypoglycosylated alpha-dystroglycan. Biochem. Biophys. Res. Commun. 2016, 477, 137–143. [Google Scholar] [CrossRef]
  90. Di Costanzo, S.; Balasubramanian, A.; Pond, H.L.; Rozkalne, A.; Pantaleoni, C.; Saredi, S.; Gupta, V.A.; Sunu, C.M.; Yu, T.W.; Kang, P.B.; et al. POMK mutations disrupt muscle development leading to a spectrum of neuromuscular presentations. Hum. Mol. Genet. 2014, 23, 5781–5792. [Google Scholar] [CrossRef]
  91. Carss, K.J.; Stevens, E.; Foley, A.R.; Cirak, S.; Riemersma, M.; Torelli, S.; Hoischen, A.; Willer, T.; van Scherpenzeel, M.; Moore, S.A.; et al. Mutations in GDP-mannose pyrophosphorylase B cause congenital and limb-girdle muscular dystrophies associated with hypoglycosylation of alpha-dystroglycan. Am. J. Hum. Genet. 2013, 93, 29–41. [Google Scholar] [CrossRef]
  92. Henderson, C.A.; Gomez, C.G.; Novak, S.M.; Mi-Mi, L.; Gregorio, C.C. Overview of the Muscle Cytoskeleton. Compr. Physiol. 2017, 7, 891–944. [Google Scholar]
  93. Capetanaki, Y.; Bloch, R.J.; Kouloumenta, A.; Mavroidis, M.; Psarras, S. Muscle intermediate filaments and their links to membranes and membranous organelles. Exp. Cell Res. 2007, 313, 2063–2076. [Google Scholar] [CrossRef]
  94. Ramspacher, C.; Steed, E.; Boselli, F.; Ferreira, R.; Faggianelli, N.; Roth, S.; Spiegelhalter, C.; Messaddeq, N.; Trinh, L.; Liebling, M.; et al. Developmental Alterations in Heart Biomechanics and Skeletal Muscle Function in Desmin Mutants Suggest an Early Pathological Root for Desminopathies. Cell Rep. 2015, 11, 1564–1576. [Google Scholar] [CrossRef]
  95. Paulin, D.; Li, Z. Desmin: A major intermediate filament protein essential for the structural integrity and function of muscle. Exp. Cell Res. 2004, 301, 1–7. [Google Scholar] [CrossRef] [PubMed]
  96. Li, M.; Andersson-Lendahl, M.; Sejersen, T.; Arner, A. Knockdown of desmin in zebrafish larvae affects interfilament spacing and mechanical properties of skeletal muscle. J. Gen. Physiol. 2013, 141, 335–345. [Google Scholar] [CrossRef] [PubMed]
  97. Kayman Kurekci, G.; Kural Mangit, E.; Koyunlar, C.; Unsal, S.; Saglam, B.; Ergin, B.; Gizer, M.; Uyanik, I.; Boustanabadimaralan Duz, N.; Korkusuz, P.; et al. Knockout of zebrafish desmin genes does not cause skeletal muscle degeneration but alters calcium flux. Sci. Rep. 2021, 11, 7505. [Google Scholar] [CrossRef] [PubMed]
  98. Nicolas, H.A.; Hua, K.; Quigley, H.; Ivare, J.; Tesson, F.; Akimenko, M.A. A CRISPR/Cas9 zebrafish lamin A/C mutant model of muscular laminopathy. Dev. Dyn. 2022, 251, 645–661. [Google Scholar] [CrossRef] [PubMed]
  99. Birely, J.; Schneider, V.A.; Santana, E.; Dosch, R.; Wagner, D.S.; Mullins, M.C.; Granato, M. Genetic screens for genes controlling motor nerve-muscle development and interactions. Dev. Biol. 2005, 280, 162–176. [Google Scholar] [CrossRef]
  100. Ruparelia, A.A.; Zhao, M.; Currie, P.D.; Bryson-Richardson, R.J. Characterization and investigation of zebrafish models of filamin-related myofibrillar myopathy. Hum. Mol. Genet. 2012, 21, 4073–4083. [Google Scholar] [CrossRef]
  101. Koshimizu, E.; Imamura, S.; Qi, J.; Toure, J.; Valdez, D.M., Jr.; Carr, C.E.; Hanai, J.; Kishi, S. Embryonic senescence and laminopathies in a progeroid zebrafish model. PLoS ONE 2011, 6, e17688. [Google Scholar] [CrossRef] [PubMed]
  102. Sarparanta, J.; Jonson, P.H.; Golzio, C.; Sandell, S.; Luque, H.; Screen, M.; McDonald, K.; Stajich, J.M.; Mahjneh, I.; Vihola, A.; et al. Mutations affecting the cytoplasmic functions of the co-chaperone DNAJB6 cause limb-girdle muscular dystrophy. Nat. Genet. 2012, 44, 450–455. [Google Scholar] [CrossRef] [PubMed]
  103. Nam, T.S.; Li, W.; Heo, S.H.; Lee, K.H.; Cho, A.; Shin, J.H.; Kim, Y.O.; Chae, J.H.; Kim, D.S.; Kim, M.K.; et al. A novel mutation in DNAJB6, p.(Phe91Leu), in childhood-onset LGMD1D with a severe phenotype. Neuromuscul. Disord. 2015, 25, 843–851. [Google Scholar] [CrossRef] [PubMed]
  104. Cheng, L.; Guo, X.F.; Yang, X.Y.; Chong, M.; Cheng, J.; Li, G.; Gui, Y.H.; Lu, D.R. Delta-sarcoglycan is necessary for early heart and muscle development in zebrafish. Biochem. Biophys. Res. Commun. 2006, 344, 1290–1299. [Google Scholar] [CrossRef] [PubMed]
  105. Guyon, J.R.; Mosley, A.N.; Jun, S.J.; Montanaro, F.; Steffen, L.S.; Zhou, Y.; Nigro, V.; Zon, L.I.; Kunkel, L.M. Delta-sarcoglycan is required for early zebrafish muscle organization. Exp. Cell Res. 2005, 304, 105–115. [Google Scholar] [CrossRef]
  106. Parton, R.G. Caveolae: Structure, Function, and Relationship to Disease. Annu. Rev. Cell Dev. Biol. 2018, 34, 111–136. [Google Scholar] [CrossRef]
  107. McNally, E.M.; de Sa Moreira, E.; Duggan, D.J.; Bonnemann, C.G.; Lisanti, M.P.; Lidov, H.G.; Vainzof, M.; Passos-Bueno, M.R.; Hoffman, E.P.; Zatz, M.; et al. Caveolin-3 in muscular dystrophy. Hum. Mol. Genet. 1998, 7, 871–877. [Google Scholar] [CrossRef]
  108. Nixon, S.J.; Wegner, J.; Ferguson, C.; Mery, P.F.; Hancock, J.F.; Currie, P.D.; Key, B.; Westerfield, M.; Parton, R.G. Zebrafish as a model for caveolin-associated muscle disease; caveolin-3 is required for myofibril organization and muscle cell patterning. Hum. Mol. Genet. 2005, 14, 1727–1743. [Google Scholar] [CrossRef]
  109. Postel, R.; Vakeel, P.; Topczewski, J.; Knoll, R.; Bakkers, J. Zebrafish integrin-linked kinase is required in skeletal muscles for strengthening the integrin-ECM adhesion complex. Dev. Biol. 2008, 318, 92–101. [Google Scholar] [CrossRef]
  110. Kawahara, G.; Serafini, P.R.; Myers, J.A.; Alexander, M.S.; Kunkel, L.M. Characterization of zebrafish dysferlin by morpholino knockdown. Biochem. Biophys. Res. Commun. 2011, 413, 358–363. [Google Scholar] [CrossRef]
  111. Roostalu, U.; Strahle, U. In vivo imaging of molecular interactions at damaged sarcolemma. Dev. Cell 2012, 22, 515–529. [Google Scholar] [CrossRef]
  112. De Oliveira, A.A.S.; Brigante, T.A.V.; Oliveira, D.P. Tail coiling assay in zebrafish (Danio rerio) embryos: Stage of development, promising positive control candidates, and selection of an appropriate organic solvent for screening of developmental neurotoxicity (DNT). Water 2021, 2, 119. [Google Scholar] [CrossRef]
  113. Ogungbemi, A.O.; Massei, R.; Altenburger, R.; Scholz, S.; Kuster, E. Assessing Combined Effects for Mixtures of Similar and Dissimilar Acting Neuroactive Substances on Zebrafish Embryo Movement. Toxics 2021, 9, 104. [Google Scholar] [CrossRef] [PubMed]
  114. Sztal, T.E.; Ruparelia, A.A.; Williams, C.; Bryson-Richardson, R.J. Using Touch-evoked Response and Locomotion Assays to Assess Muscle Performance and Function in Zebrafish. J. Vis. Exp. 2016, 116, e54431. [Google Scholar]
  115. Brustein, E.; Saint-Amant, L.; Buss, R.R.; Chong, M.; McDearmid, J.R.; Drapeau, P. Steps during the development of the zebrafish locomotor network. J. Physiol. Paris 2003, 97, 77–86. [Google Scholar] [CrossRef] [PubMed]
  116. Drapeau, P.; Saint-Amant, L.; Buss, R.R.; Chong, M.; McDearmid, J.R.; Brustein, E. Development of the locomotor network in zebrafish. Prog. Neurobiol. 2002, 68, 85–111. [Google Scholar] [CrossRef] [PubMed]
  117. Schiavone, M.; Zulian, A.; Menazza, S.; Petronilli, V.; Argenton, F.; Merlini, L.; Sabatelli, P.; Bernardi, P. Alisporivir rescues defective mitochondrial respiration in Duchenne muscular dystrophy. Pharmacol. Res. 2017, 125 Pt B, 122–131. [Google Scholar] [CrossRef]
  118. Stocco, A.; Smolina, N.; Sabatelli, P.; Sileikyte, J.; Artusi, E.; Mouly, V.; Cohen, M.; Forte, M.; Schiavone, M.; Bernardi, P. Treatment with a triazole inhibitor of the mitochondrial permeability transition pore fully corrects the pathology of sapje zebrafish lacking dystrophin. Pharmacol. Res. 2021, 165, 105421. [Google Scholar] [CrossRef]
  119. Zulian, A.; Rizzo, E.; Schiavone, M.; Palma, E.; Tagliavini, F.; Blaauw, B.; Merlini, L.; Maraldi, N.M.; Sabatelli, P.; Braghetta, P.; et al. NIM811, a cyclophilin inhibitor without immunosuppressive activity, is beneficial in collagen VI congenital muscular dystrophy models. Hum. Mol. Genet. 2014, 23, 5353–5363. [Google Scholar] [CrossRef]
  120. Mrinalini, R.; Tamilanban, T.; Naveen Kumar, V.; Manasa, K. Zebrafish—The Neurobehavioural Model in Trend. Neuroscience 2022, in press. [Google Scholar] [CrossRef]
  121. Ahmad, F.; Noldus, L.P.; Tegelenbosch, R.A.; Richardson, M.K. Zebrafish embryos and larvae in behavioural assays. Behaviour 2012, 149, 1241–1281. [Google Scholar]
  122. Orger, M.B.; de Polavieja, G.G. Zebrafish Behavior: Opportunities and Challenges. Annu. Rev. Neurosci. 2017, 40, 125–147. [Google Scholar] [CrossRef] [PubMed]
  123. Rosa, J.G.S.; Lima, C.; Lopes-Ferreira, M. Zebrafish Larvae Behavior Models as a Tool for Drug Screenings and Pre-Clinical Trials: A Review. Int. J. Mol. Sci. 2022, 23, 6647. [Google Scholar] [CrossRef] [PubMed]
  124. Huang, S.H.; Tsao, C.W.; Fang, Y.H. A Miniature Intermittent-Flow Respirometry System with a 3D-Printed, Palm-Sized Zebrafish Treadmill for Measuring Rest and Activity Metabolic Rates. Sensors 2020, 20, 5088. [Google Scholar] [CrossRef] [PubMed]
  125. Masse, A.J.; Thomas, J.K.; Janz, D.M. Reduced swim performance and aerobic capacity in adult zebrafish exposed to waterborne selenite. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2013, 157, 266–271. [Google Scholar] [CrossRef]
  126. Palstra, A.P.; Kals, J.; Bohm, T.; Bastiaansen, J.W.M.; Komen, H. Swimming Performance and Oxygen Consumption as Non-lethal Indicators of Production Traits in Atlantic Salmon and Gilthead Seabream. Front. Physiol. 2020, 11, 759. [Google Scholar] [CrossRef]
  127. Wakamatsu, Y.; Ogino, K.; Hirata, H. Swimming capability of zebrafish is governed by water temperature, caudal fin length and genetic background. Sci. Rep. 2019, 9, 16307. [Google Scholar] [CrossRef]
  128. Cresci, A.; De Rosa, R.; Agnisola, C. Assessing the Influence of Personality on Sensitivity to Magnetic Fields in Zebrafish. Available online: https://app.jove.com/it/v/59229/assessing-the-influence-of-personality-on-sensitivity-to-magnetic-fields-in-zebrafish (accessed on 27 April 2023).
  129. Sloboda, D.D.; Claflin, D.R.; Dowling, J.J.; Brooks, S.V. Force measurement during contraction to assess muscle function in zebrafish larvae. J. Vis. Exp. 2013, e50539. [Google Scholar] [CrossRef]
  130. Widrick, J.J.; Alexander, M.S.; Sanchez, B.; Gibbs, D.E.; Kawahara, G.; Beggs, A.H.; Kunkel, L.M. Muscle dysfunction in a zebrafish model of Duchenne muscular dystrophy. Physiol. Genom. 2016, 48, 850–860. [Google Scholar] [CrossRef]
  131. Dowling, J.J.; Vreede, A.P.; Low, S.E.; Gibbs, E.M.; Kuwada, J.Y.; Bonnemann, C.G.; Feldman, E.L. Loss of myotubularin function results in T-tubule disorganization in zebrafish and human myotubular myopathy. PLoS Genet. 2009, 5, e1000372. [Google Scholar] [CrossRef]
  132. Telfer, W.R.; Nelson, D.D.; Waugh, T.; Brooks, S.V.; Dowling, J.J. Neb: A zebrafish model of nemaline myopathy due to nebulin mutation. Dis. Model Mech. 2012, 5, 389–396. [Google Scholar] [CrossRef] [PubMed]
  133. Hirata, H.; Saint-Amant, L.; Waterbury, J.; Cui, W.; Zhou, W.; Li, Q.; Goldman, D.; Granato, M.; Kuwada, J.Y. accordion, a zebrafish behavioral mutant, has a muscle relaxation defect due to a mutation in the ATPase Ca2+ pump SERCA1. Development 2004, 131, 5457–5468. [Google Scholar] [CrossRef] [PubMed]
  134. Smith, L.L.; Beggs, A.H.; Gupta, V.A. Analysis of Skeletal Muscle Defects in Larval Zebrafish by Birefringence and Touch-Evoke Escape Response Assays. Available online: https://app.jove.com/it/v/50925/analysis-of-skeletal-muscle-defects-in-larval-zebrafish-by-birefringence-and-touch-evoke-escape-response-assays (accessed on 27 April 2023).
  135. Berger, J.; Sztal, T.; Currie, P.D. Quantification of birefringence readily measures the level of muscle damage in zebrafish. Biochem. Biophys. Res. Commun. 2012, 423, 785–788. [Google Scholar] [CrossRef] [PubMed]
  136. Berger, J.; Li, M.; Berger, S.; Meilak, M.; Rientjes, J.; Currie, P.D. Effect of Ataluren on dystrophin mutations. J. Cell. Mol. Med. 2020, 24, 6680–6689. [Google Scholar] [CrossRef] [PubMed]
  137. Kawahara, G.; Karpf, J.A.; Myers, J.A.; Alexander, M.S.; Guyon, J.R.; Kunkel, L.M. Drug screening in a zebrafish model of Duchenne muscular dystrophy. Proc. Natl. Acad. Sci. USA 2011, 108, 5331–5336. [Google Scholar] [CrossRef] [PubMed]
  138. Moro, E.; Vettori, A.; Porazzi, P.; Schiavone, M.; Rampazzo, E.; Casari, A.; Ek, O.; Facchinello, N.; Astone, M.; Zancan, I.; et al. Generation and application of signaling pathway reporter lines in zebrafish. Mol. Genet. Genom. 2013, 288, 231–242. [Google Scholar] [CrossRef]
  139. Goody, M.; Henry, C. Phalloidin Staining and Immunohistochemistry of Zebrafish Embryos. Bio-Protocol 2013, 3, e786. [Google Scholar] [CrossRef]
  140. Zou, Y.; Wang, A.; Huang, L.; Zhu, X.; Hu, Q.; Zhang, Y.; Chen, X.; Li, F.; Wang, Q.; Wang, H.; et al. Illuminating NAD(+) Metabolism in Live Cells and In Vivo Using a Genetically Encoded Fluorescent Sensor. Dev. Cell 2020, 53, 240–252.e7. [Google Scholar] [CrossRef]
  141. Zou, Y.; Wang, A.; Shi, M.; Chen, X.; Liu, R.; Li, T.; Zhang, C.; Zhang, Z.; Zhu, L.; Ju, Z.; et al. Analysis of redox landscapes and dynamics in living cells and in vivo using genetically encoded fluorescent sensors. Nat. Protoc. 2018, 13, 2362–2386. [Google Scholar] [CrossRef]
  142. Scott, C.A.; Carney, T.J.; Amaya, E. Aerobic glycolysis is important for zebrafish larval wound closure and tail regeneration. Wound Repair Regen. 2022, 30, 665–680. [Google Scholar] [CrossRef]
  143. Panieri, E.; Santoro, M.M. Data on metabolic-dependent antioxidant response in the cardiovascular tissues of living zebrafish under stress conditions. Data Brief 2017, 12, 427–432. [Google Scholar] [CrossRef] [PubMed]
  144. Lackmann, C.; Santos, M.M.; Rainieri, S.; Barranco, A.; Hollert, H.; Spirhanzlova, P.; Velki, M.; Seiler, T.B. Novel procedures for whole organism detection and quantification of fluorescence as a measurement for oxidative stress in zebrafish (Danio rerio) larvae. Chemosphere 2018, 197, 200–209. [Google Scholar] [CrossRef] [PubMed]
  145. Mugoni, V.; Camporeale, A.; Santoro, M.M. Analysis of oxidative stress in zebrafish embryos. J. Vis. Exp. 2014, 51328. [Google Scholar] [CrossRef]
  146. Jackson, H.E.; Ono, Y.; Wang, X.; Elworthy, S.; Cunliffe, V.T.; Ingham, P.W. The role of Sox6 in zebrafish muscle fiber type specification. Skelet. Muscle 2015, 5, 2. [Google Scholar] [CrossRef]
  147. Kettunen, P. Calcium Imaging in the Zebrafish. Adv. Exp. Med. Biol. 2020, 1131, 901–942. [Google Scholar]
  148. Okamoto, S.I.; Hatta, K. Ca2+-imaging and photo-manipulation of the simple gut of zebrafish larvae in vivo. Sci. Rep. 2022, 12, 2018. [Google Scholar] [CrossRef]
  149. Tsuruwaka, Y.; Shimada, E.; Tsutsui, K.; Ogawa, T. Ca2+ dynamics in zebrafish morphogenesis. PeerJ 2017, 5, e2894. [Google Scholar] [CrossRef]
  150. Vanwalleghem, G.; Constantin, L.; Scott, E.K. Calcium Imaging and the Curse of Negativity. Front. Neural. Circuits 2020, 14, 607391. [Google Scholar] [CrossRef]
  151. Kole, K.; Voesenek, B.J.B.; Brinia, M.E.; Petersen, N.; Kole, M.H.P. Parvalbumin basket cell myelination accumulates axonal mitochondria to internodes. Nat. Commun. 2022, 13, 7598. [Google Scholar] [CrossRef]
  152. Bond, S.T.; McEwen, K.A.; Yoganantharajah, P.; Gibert, Y. Live Metabolic Profile Analysis of Zebrafish Embryos Using a Seahorse XF 24 Extracellular Flux Analyzer. Methods Mol. Biol. 2018, 1797, 393–401. [Google Scholar]
  153. Housley, M.P.; Njaine, B.; Ricciardi, F.; Stone, O.A.; Holper, S.; Kruger, M.; Kostin, S.; Stainier, D.Y. Cavin4b/Murcb Is Required for Skeletal Muscle Development and Function in Zebrafish. PLoS Genet. 2016, 12, e1006099. [Google Scholar] [CrossRef] [PubMed]
  154. Spreafico, M.; Cafora, M.; Bragato, C.; Capitanio, D.; Marasca, F.; Bodega, B.; De Palma, C.; Mora, M.; Gelfi, C.; Marozzi, A.; et al. Targeting HDAC8 to ameliorate skeletal muscle differentiation in Duchenne muscular dystrophy. Pharmacol. Res. 2021, 170, 105750. [Google Scholar] [CrossRef] [PubMed]
  155. Yamakawa, N.; Vanbeselaere, J.; Chang, L.Y.; Yu, S.Y.; Ducrocq, L.; Harduin-Lepers, A.; Kurata, J.; Aoki-Kinoshita, K.F.; Sato, C.; Khoo, K.H.; et al. Systems glycomics of adult zebrafish identifies organ-specific sialylation and glycosylation patterns. Nat. Commun. 2018, 9, 4647. [Google Scholar] [CrossRef] [PubMed]
  156. Vettori, A.; Greenald, D.; Wilson, G.K.; Peron, M.; Facchinello, N.; Markham, E.; Sinnakaruppan, M.; Matthews, L.C.; McKeating, J.A.; Argenton, F.; et al. Glucocorticoids promote Von Hippel Lindau degradation and Hif-1alpha stabilization. Proc. Natl. Acad. Sci. USA 2017, 114, 9948–9953. [Google Scholar] [CrossRef]
  157. Barresi, M.J.; Stickney, H.L.; Devoto, S.H. The zebrafish slow-muscle-omitted gene product is required for Hedgehog signal transduction and the development of slow muscle identity. Development 2000, 127, 2189–2199. [Google Scholar] [CrossRef]
  158. Sasaki, H.; Hui, C.; Nakafuku, M.; Kondoh, H. A binding site for Gli proteins is essential for HNF-3beta floor plate enhancer activity in transgenics and can respond to Shh in vitro. Development 1997, 124, 1313–1322. [Google Scholar] [CrossRef]
  159. Corallo, D.; Schiavinato, A.; Trapani, V.; Moro, E.; Argenton, F.; Bonaldo, P. Emilin3 is required for notochord sheath integrity and interacts with Scube2 to regulate notochord-derived Hedgehog signals. Development 2013, 140, 4594–4601. [Google Scholar] [CrossRef]
  160. Mich, J.K.; Payumo, A.Y.; Rack, P.G.; Chen, J.K. In vivo imaging of Hedgehog pathway activation with a nuclear fluorescent reporter. PLoS ONE 2014, 9, e103661. [Google Scholar] [CrossRef]
  161. Barolo, S. Transgenic Wnt/TCF pathway reporters: All you need is Lef? Oncogene 2006, 25, 7505–7511. [Google Scholar] [CrossRef]
  162. Dorsky, R.I.; Sheldahl, L.C.; Moon, R.T. A transgenic Lef1/beta-catenin-dependent reporter is expressed in spatially restricted domains throughout zebrafish development. Dev. Biol. 2002, 241, 229–237. [Google Scholar] [CrossRef]
  163. Moro, E.; Ozhan-Kizil, G.; Mongera, A.; Beis, D.; Wierzbicki, C.; Young, R.M.; Bournele, D.; Domenichini, A.; Valdivia, L.E.; Lum, L.; et al. In vivo Wnt signaling tracing through a transgenic biosensor fish reveals novel activity domains. Dev. Biol. 2012, 366, 327–340. [Google Scholar] [CrossRef] [PubMed]
  164. Pyati, U.J.; Webb, A.E.; Kimelman, D. Transgenic zebrafish reveal stage-specific roles for Bmp signaling in ventral and posterior mesoderm development. Development 2005, 132, 2333–2343. [Google Scholar] [CrossRef]
  165. Casari, A.; Schiavone, M.; Facchinello, N.; Vettori, A.; Meyer, D.; Tiso, N.; Moro, E.; Argenton, F. A Smad3 transgenic reporter reveals TGF-beta control of zebrafish spinal cord development. Dev. Biol. 2014, 396, 81–93. [Google Scholar] [CrossRef]
  166. Collery, R.F.; Link, B.A. Dynamic smad-mediated BMP signaling revealed through transgenic zebrafish. Dev. Dyn. 2011, 240, 712–722. [Google Scholar] [CrossRef] [PubMed]
  167. Ramel, M.C.; Hill, C.S. The ventral to dorsal BMP activity gradient in the early zebrafish embryo is determined by graded expression of BMP ligands. Dev. Biol. 2013, 378, 170–182. [Google Scholar] [CrossRef] [PubMed]
  168. Wu, M.Y.; Ramel, M.C.; Howell, M.; Hill, C.S. SNW1 is a critical regulator of spatial BMP activity, neural plate border formation, and neural crest specification in vertebrate embryos. PLoS Biol. 2011, 9, e1000593. [Google Scholar] [CrossRef] [PubMed]
  169. Giuliodori, A.; Beffagna, G.; Marchetto, G.; Fornetto, C.; Vanzi, F.; Toppo, S.; Facchinello, N.; Santimaria, M.; Vettori, A.; Rizzo, S.; et al. Loss of cardiac Wnt/beta-catenin signalling in desmoplakin-deficient AC8 zebrafish models is rescuable by genetic and pharmacological intervention. Cardiovasc. Res. 2018, 114, 1082–1097. [Google Scholar] [CrossRef]
  170. Merkulov, V.M.; Merkulova, T.I. Structural variants of glucocorticoid receptor binding sites and different versions of positive glucocorticoid responsive elements: Analysis of GR-TRRD database. J. Steroid Biochem. Mol. Biol. 2009, 115, 1–8. [Google Scholar] [CrossRef]
  171. Benato, F.; Colletti, E.; Skobo, T.; Moro, E.; Colombo, L.; Argenton, F.; Dalla Valle, L. A living biosensor model to dynamically trace glucocorticoid transcriptional activity during development and adult life in zebrafish. Mol. Cell. Endocrinol. 2014, 392, 60–72. [Google Scholar] [CrossRef]
  172. Wang, X.J.; Hayes, J.D.; Wolf, C.R. Generation of a stable antioxidant response element-driven reporter gene cell line and its use to show redox-dependent activation of nrf2 by cancer chemotherapeutic agents. Cancer Res. 2006, 66, 10983–10994. [Google Scholar] [CrossRef]
  173. Badenetti, L.; Manzoli, R.; Rubin, M.; Cozza, G.; Moro, E. Monitoring Nrf2/ARE Pathway Activity with a New Zebrafish Reporter System. Int. J. Mol. Sci. 2023, 24, 6804. [Google Scholar] [CrossRef] [PubMed]
  174. Nguyen, P.D.; Gurevich, D.B.; Sonntag, C.; Hersey, L.; Alaei, S.; Nim, H.T.; Siegel, A.; Hall, T.E.; Rossello, F.J.; Boyd, S.E.; et al. Muscle Stem Cells Undergo Extensive Clonal Drift during Tissue Growth via Meox1-Mediated Induction of G2 Cell-Cycle Arrest. Cell Stem Cell 2017, 21, 107–119.e6. [Google Scholar] [CrossRef] [PubMed]
  175. Seger, C.; Hargrave, M.; Wang, X.; Chai, R.J.; Elworthy, S.; Ingham, P.W. Analysis of Pax7 expressing myogenic cells in zebrafish muscle development, injury, and models of disease. Dev. Dyn. 2011, 240, 2440–2451. [Google Scholar] [CrossRef] [PubMed]
  176. Ellingsen, S.; Laplante, M.A.; Konig, M.; Kikuta, H.; Furmanek, T.; Hoivik, E.A.; Becker, T.S. Large-scale enhancer detection in the zebrafish genome. Development 2005, 132, 3799–3811. [Google Scholar] [CrossRef]
  177. Ingham, P.W.; Kim, H.R. Hedgehog signalling and the specification of muscle cell identity in the zebrafish embryo. Exp. Cell Res. 2005, 306, 336–342. [Google Scholar] [CrossRef]
  178. Maurya, A.K.; Tan, H.; Souren, M.; Wang, X.; Wittbrodt, J.; Ingham, P.W. Integration of Hedgehog and BMP signalling by the engrailed2a gene in the zebrafish myotome. Development 2011, 138, 755–765. [Google Scholar] [CrossRef]
  179. Nguyen, P.D.; Currie, P.D. Guidelines and best practices in successfully using Zebrabow for lineage tracing multiple cells within tissues. Methods 2018, 150, 63–67. [Google Scholar] [CrossRef]
  180. Gurevich, D.B.; Nguyen, P.D.; Siegel, A.L.; Ehrlich, O.V.; Sonntag, C.; Phan, J.M.; Berger, S.; Ratnayake, D.; Hersey, L.; Berger, J.; et al. Asymmetric division of clonal muscle stem cells coordinates muscle regeneration in vivo. Science 2016, 353, aad9969. [Google Scholar] [CrossRef]
  181. Gong, Z.; Wan, H.; Tay, T.L.; Wang, H.; Chen, M.; Yan, T. Development of transgenic fish for ornamental and bioreactor by strong expression of fluorescent proteins in the skeletal muscle. Biochem. Biophys. Res. Commun. 2003, 308, 58–63. [Google Scholar] [CrossRef]
  182. Ju, B.; Chong, S.W.; He, J.; Wang, X.; Xu, Y.; Wan, H.; Tong, Y.; Yan, T.; Korzh, V.; Gong, Z. Recapitulation of fast skeletal muscle development in zebrafish by transgenic expression of GFP under the mylz2 promoter. Dev. Dyn. 2003, 227, 14–26. [Google Scholar] [CrossRef]
  183. Moore, C.A.; Parkin, C.A.; Bidet, Y.; Ingham, P.W. A role for the Myoblast city homologues Dock1 and Dock5 and the adaptor proteins Crk and Crk-like in zebrafish myoblast fusion. Development 2007, 134, 3145–3153. [Google Scholar] [CrossRef]
  184. Berger, J.; Tarakci, H.; Berger, S.; Li, M.; Hall, T.E.; Arner, A.; Currie, P.D. Loss of Tropomodulin4 in the zebrafish mutant trage causes cytoplasmic rod formation and muscle weakness reminiscent of nemaline myopathy. Dis. Model Mech. 2014, 7, 1407–1415. [Google Scholar] [CrossRef] [PubMed]
  185. Higashijima, S.; Okamoto, H.; Ueno, N.; Hotta, Y.; Eguchi, G. High-frequency generation of transgenic zebrafish which reliably express GFP in whole muscles or the whole body by using promoters of zebrafish origin. Dev. Biol. 1997, 192, 289–299. [Google Scholar] [CrossRef]
  186. Meneghetti, G.; Skobo, T.; Chrisam, M.; Facchinello, N.; Fontana, C.M.; Bellesso, S.; Sabatelli, P.; Raggi, F.; Cecconi, F.; Bonaldo, P.; et al. The epg5 knockout zebrafish line: A model to study Vici syndrome. Autophagy 2019, 15, 1438–1454. [Google Scholar] [CrossRef]
  187. Ellis, K.; Bagwell, J.; Bagnat, M. Notochord vacuoles are lysosome-related organelles that function in axis and spine morphogenesis. J. Cell Biol. 2013, 200, 667–679. [Google Scholar] [CrossRef] [PubMed]
  188. Clark, B.S.; Winter, M.; Cohen, A.R.; Link, B.A. Generation of Rab-based transgenic lines for in vivo studies of endosome biology in zebrafish. Dev. Dyn. 2011, 240, 2452–2465. [Google Scholar] [CrossRef] [PubMed]
  189. Zhang, M.; Sun, T.; Jian, C.; Lei, L.; Han, P.; Lv, Q.; Yang, R.; Zhou, X.; Xu, J.; Hu, Y.; et al. Remodeling of Mitochondrial Flashes in Muscular Development and Dystrophy in Zebrafish. PLoS ONE 2015, 10, e0132567. [Google Scholar] [CrossRef]
  190. Kim, M.J.; Kang, K.H.; Kim, C.H.; Choi, S.Y. Real-time imaging of mitochondria in transgenic zebrafish expressing mitochondrially targeted GFP. Biotechniques 2008, 45, 331–334. [Google Scholar]
  191. Facchinello, N.; Laquatra, C.; Locatello, L.; Beffagna, G.; Branas Casas, R.; Fornetto, C.; Dinarello, A.; Martorano, L.; Vettori, A.; Risato, G.; et al. Efficient clofilium tosylate-mediated rescue of POLG-related disease phenotypes in zebrafish. Cell Death Dis. 2021, 12, 100. [Google Scholar] [CrossRef]
  192. Chen, Y.H.; Wang, Y.H.; Chang, M.Y.; Lin, C.Y.; Weng, C.W.; Westerfield, M.; Tsai, H.J. Multiple upstream modules regulate zebrafish myf5 expression. BMC Dev. Biol. 2007, 7, 1. [Google Scholar] [CrossRef]
  193. Elworthy, S.; Hargrave, M.; Knight, R.; Mebus, K.; Ingham, P.W. Expression of multiple slow myosin heavy chain genes reveals a diversity of zebrafish slow twitch muscle fibres with differing requirements for Hedgehog and Prdm1 activity. Development 2008, 135, 2115–2126. [Google Scholar] [CrossRef] [PubMed]
  194. Asaduzzaman, M.; Kinoshita, S.; Bhuiyan, S.S.; Asakawa, S.; Watabe, S. Stimulatory and inhibitory mechanisms of slow muscle-specific myosin heavy chain gene expression in fish: Transient and transgenic analysis of torafugu MYH(M86-2) promoter in zebrafish embryos. Exp. Cell Res. 2013, 319, 820–837. [Google Scholar] [CrossRef] [PubMed]
  195. Talbot, J.C.; Teets, E.M.; Ratnayake, D.; Duy, P.Q.; Currie, P.D.; Amacher, S.L. Muscle precursor cell movements in zebrafish are dynamic and require Six family genes. Development 2019, 146, dev171421. [Google Scholar] [CrossRef] [PubMed]
  196. Nguyen, P.D.; Hollway, G.E.; Sonntag, C.; Miles, L.B.; Hall, T.E.; Berger, S.; Fernandez, K.J.; Gurevich, D.B.; Cole, N.J.; Alaei, S.; et al. Haematopoietic stem cell induction by somite-derived endothelial cells controlled by meox1. Nature 2014, 512, 314–318. [Google Scholar] [CrossRef] [PubMed]
  197. Pan, Y.A.; Freundlich, T.; Weissman, T.A.; Schoppik, D.; Wang, X.C.; Zimmerman, S.; Ciruna, B.; Sanes, J.R.; Lichtman, J.W.; Schier, A.F. Zebrabow: Multispectral cell labeling for cell tracing and lineage analysis in zebrafish. Development 2013, 140, 2835–2846. [Google Scholar] [CrossRef]
  198. Wasala, N.B.; Chen, S.J.; Duan, D. Duchenne muscular dystrophy animal models for high-throughput drug discovery and precision medicine. Expert Opin. Drug Discov. 2020, 15, 443–456. [Google Scholar] [CrossRef] [PubMed]
  199. Sicinski, P.; Geng, Y.; Ryder-Cook, A.S.; Barnard, E.A.; Darlison, M.G.; Barnard, P.J. The molecular basis of muscular dystrophy in the mdx mouse: A point mutation. Science 1989, 244, 1578–1580. [Google Scholar] [CrossRef]
  200. Calyjur, P.C.; Almeida Cde, F.; Ayub-Guerrieri, D.; Ribeiro, A.F., Jr.; Fernandes Sde, A.; Ishiba, R.; Santos, A.L.; Onofre-Oliveira, P.; Vainzof, M. The mdx Mutation in the 129/Sv Background Results in a Milder Phenotype: Transcriptome Comparative Analysis Searching for the Protective Factors. PLoS ONE 2016, 11, e0150748. [Google Scholar] [CrossRef]
  201. Coley, W.D.; Bogdanik, L.; Vila, M.C.; Yu, Q.; Van Der Meulen, J.H.; Rayavarapu, S.; Novak, J.S.; Nearing, M.; Quinn, J.L.; Saunders, A.; et al. Effect of genetic background on the dystrophic phenotype in mdx mice. Hum. Mol. Genet. 2016, 25, 130–145. [Google Scholar] [CrossRef]
  202. Hammers, D.W.; Hart, C.C.; Matheny, M.K.; Wright, L.A.; Armellini, M.; Barton, E.R.; Sweeney, H.L. The D2.mdx mouse as a preclinical model of the skeletal muscle pathology associated with Duchenne muscular dystrophy. Sci. Rep. 2020, 10, 14070. [Google Scholar] [CrossRef]
  203. Zon, L.I.; Peterson, R.T. In vivo drug discovery in the zebrafish. Nat. Rev. Drug Discov. 2005, 4, 35–44. [Google Scholar] [CrossRef] [PubMed]
  204. Licitra, R.; Marchese, M.; Brogi, L.; Fronte, B.; Pitto, L.; Santorelli, F.M. Nutraceutical Screening in a Zebrafish Model of Muscular Dystrophy: Gingerol as a Possible Food Aid. Nutrients 2021, 13, 998. [Google Scholar] [CrossRef] [PubMed]
  205. Waugh, T.A.; Horstick, E.; Hur, J.; Jackson, S.W.; Davidson, A.E.; Li, X.; Dowling, J.J. Fluoxetine prevents dystrophic changes in a zebrafish model of Duchenne muscular dystrophy. Hum. Mol. Genet. 2014, 23, 4651–4662. [Google Scholar] [CrossRef] [PubMed]
  206. Farr, G.H., 3rd; Morris, M.; Gomez, A.; Pham, T.; Kilroy, E.; Parker, E.U.; Said, S.; Henry, C.; Maves, L. A novel chemical-combination screen in zebrafish identifies epigenetic small molecule candidates for the treatment of Duchenne muscular dystrophy. Skelet. Muscle 2020, 10, 29. [Google Scholar] [CrossRef] [PubMed]
  207. Spinazzola, J.M.; Lambert, M.R.; Gibbs, D.E.; Conner, J.R.; Krikorian, G.L.; Pareek, P.; Rago, C.; Kunkel, L.M. Effect of serotonin modulation on dystrophin-deficient zebrafish. Biol. Open 2020, 9, bio53363. [Google Scholar] [CrossRef]
  208. Lambert, M.R.; Spinazzola, J.M.; Widrick, J.J.; Pakula, A.; Conner, J.R.; Chin, J.E.; Owens, J.M.; Kunkel, L.M. PDE10A Inhibition Reduces the Manifestation of Pathology in DMD Zebrafish and Represses the Genetic Modifier PITPNA. Mol. Ther. 2021, 29, 1086–1101. [Google Scholar] [CrossRef]
  209. Ketley, A.; Chen, C.Z.; Li, X.; Arya, S.; Robinson, T.E.; Granados-Riveron, J.; Udosen, I.; Morris, G.E.; Holt, I.; Furling, D.; et al. High-content screening identifies small molecules that remove nuclear foci, affect MBNL distribution and CELF1 protein levels via a PKC-independent pathway in myotonic dystrophy cell lines. Hum. Mol. Genet. 2014, 23, 1551–1562. [Google Scholar] [CrossRef]
  210. Roy, S.; Sileikyte, J.; Schiavone, M.; Neuenswander, B.; Argenton, F.; Aube, J.; Hedrick, M.P.; Chung, T.D.; Forte, M.A.; Bernardi, P.; et al. Discovery, Synthesis, and Optimization of Diarylisoxazole-3-carboxamides as Potent Inhibitors of the Mitochondrial Permeability Transition Pore. ChemMedChem 2015, 10, 1655–1671. [Google Scholar] [CrossRef]
  211. Sileikyte, J.; Devereaux, J.; de Jong, J.; Schiavone, M.; Jones, K.; Nilsen, A.; Bernardi, P.; Forte, M.; Cohen, M.S. Second-Generation Inhibitors of the Mitochondrial Permeability Transition Pore with Improved Plasma Stability. ChemMedChem 2019, 14, 1771–1782. [Google Scholar] [CrossRef]
Figure 1. Zebrafish muscle organization. As teleost fish, zebrafish are characterized by two distinct types of muscle fibers: slow-red (in red) and fast-white muscle (in gray) (panel C). Regardless of the type, zebrafish skeletal muscle fibers are organized into myofibrils, which are long protein bundles surrounded by the so-called sarcoplasm and held together by the sarcolemma (panel B). Several proteins are involved in the functioning and maintenance of myofibril integrity. The (panel A) shows a model of the spatial configuration of the Dystrophin-associated glycoprotein complex (DGC) and related proteins that provide the connection of the sarcolemma to the basal lamina. In this panel is shown the dystrophin core complex (dystrophin, dystroglycans, sarcoglycans, sarcospan, syntrophins and dystrobrevins) as well as other DCG-associated proteins that forms a network with the extracellular matrix, the cytoskeleton, the sarcolemma, and the sarcomere. Because of graphical limitations, the physiological proportions of both cytoskeletal filaments and single proteins are not represented here. MD-associated genes encoding the (i) basal membrane and extracellular matrix proteins (ii) cytosolic proteins, (iii) dystroglycan and α-DG glycosylation-related proteins, (iv) nuclear envelope and cytoskeleton proteins, and (v) membrane proteins are listed in Tables reported below in the Chapter 2 of this review. Abbreviation: IFs, intermediate filaments. Image was partially created with BioRender (www.biorender.com), accessed on 2 March 2023.
Figure 1. Zebrafish muscle organization. As teleost fish, zebrafish are characterized by two distinct types of muscle fibers: slow-red (in red) and fast-white muscle (in gray) (panel C). Regardless of the type, zebrafish skeletal muscle fibers are organized into myofibrils, which are long protein bundles surrounded by the so-called sarcoplasm and held together by the sarcolemma (panel B). Several proteins are involved in the functioning and maintenance of myofibril integrity. The (panel A) shows a model of the spatial configuration of the Dystrophin-associated glycoprotein complex (DGC) and related proteins that provide the connection of the sarcolemma to the basal lamina. In this panel is shown the dystrophin core complex (dystrophin, dystroglycans, sarcoglycans, sarcospan, syntrophins and dystrobrevins) as well as other DCG-associated proteins that forms a network with the extracellular matrix, the cytoskeleton, the sarcolemma, and the sarcomere. Because of graphical limitations, the physiological proportions of both cytoskeletal filaments and single proteins are not represented here. MD-associated genes encoding the (i) basal membrane and extracellular matrix proteins (ii) cytosolic proteins, (iii) dystroglycan and α-DG glycosylation-related proteins, (iv) nuclear envelope and cytoskeleton proteins, and (v) membrane proteins are listed in Tables reported below in the Chapter 2 of this review. Abbreviation: IFs, intermediate filaments. Image was partially created with BioRender (www.biorender.com), accessed on 2 March 2023.
Ijms 24 08314 g001
Table 1. Zebrafish models for MDs linked to the basal membrane and ECM proteins.
Table 1. Zebrafish models for MDs linked to the basal membrane and ECM proteins.
Gene SymbolProteinTechniqueRef.
lama2Laminin-α2ENU screen
ENU screen
[8]
[38]
col6a1Collagen, type VI, alpha 1CRISPR/Cas9
TALEN mut.
MO
[44]
[45]
[46]
col6a3Collagen, type VI, alpha 3MO[46]
col22a1Collagen, type XXII, alpha 1MO
CRISPR/Cas9
[43]
[47]
Table 2. Zebrafish models for MDs linked to cytosolic proteins.
Table 2. Zebrafish models for MDs linked to cytosolic proteins.
Gene SymbolProteinTechniqueRef.
dmdDystrophinGenetic screen
ENU screen
MO
[49]
[52]
[62]
dux4Double homeobox protein 4mRNA injection
Tol2 transposon system
Tol2 transposon system
[53]
[54]
[55]
cavin1a/bCaveolae-associated protein 1MO[63]
hnrnpdlHeterogeneous nuclear ribonucleoprotein D-like MO[64]
mbnl1/2/3Muscleblind-like splicing regulatorCRISPR/Cas9
mRNA injection
[56]
[58]
vcp Valosin containing protein CRISPR/Cas9
MO
[59]
[60]
[61]
mbnl2Muscleblind-like splicing regulator 2MO[57]
tcapTelethoninMO[65]
ttnTitinMO[66]
bvesBlood vessel epicardial substanceMO; TALEN mut.[67]
inpp5ka/b Inositol polyphosphate 5-phosphatase K MO[68,69]
craa/b Cristallin MO[61]
bag3 BCL2 associated athanogene 3 MO[61,70]
fhl1a/b Four and a half LIM domains 1 MO[61,71]
ldb3 LIM domain binding 3 MO[61,72]
Table 3. Zebrafish models for MDs linked to dystroglycan and α-DG glycosylation-related proteins.
Table 3. Zebrafish models for MDs linked to dystroglycan and α-DG glycosylation-related proteins.
Gene SymbolProteinTechniqueRef.
Primary dystroglycanopathy
dag1DystroglycanMO
ENU screen
[48]
[75]
Secondary dystroglycanopathies
pomt1O-mannosyl transferaseMO[78]
pomt2O-mannosyl transferaseMO[78]
pomgnt2O-mannose β-1,4-NacetylglucosaminyltransferaseMO[79]
fktnFukutinMO
MO
[80]
[81]
fkrpFukutin-related proteinTALEN mut.; Tol2 transposon system
Zinc finger nuclease KO; CRISPR/Cas9
MO
MO
MO
MO
[76]
[77]
[80]
[81]
[82]
[83]
ispdCDP-ribitol pyrophosphorylaseMO[84]
rxylt1β−1,4-xylosyl transferaseMO[85]
b3galnt2β−1,3-N-acetylgalactosaminyltransferaseMO[86]
b4gat1β−1,4-glucuronyltransferaseMO[87]
dpm1Dolichol–phosphate mannose synthaseMO[88,89]
dpm2Dolichol–phosphate mannose synthaseMO[89]
dpm3Dolichol–phosphate mannose synthaseMO[89]
pomkO-mannose kinaseMO[90]
gmppbGDP-mannose pyrophosphorylaseMO[91]
Table 4. Zebrafish models for MDs linked to the nuclear envelope and cytoskeleton proteins.
Table 4. Zebrafish models for MDs linked to the nuclear envelope and cytoskeleton proteins.
Gene SymbolProteinTechniqueRef.
desma/bDesminMO
ENU screen
MO
[61]
[94]
[96]
lmnaLamin-A/CCRISPR/Cas9 mut.
MO
[98]
[101]
myot Myotilin MO[61]
dnajb6a/bdnaJ homolog subfamily B member 6MO
MO
mRNA injection
[61]
[102]
[103]
pleca/b Plectin MO[61]
flnca/b Filamin C ENU screen, MO;
Genetic screen
[99]
[100]
sgcdδ-sarcoglycanMO[104,105]
dmpkMyotonic dystrophy protein kinasemRNA injection[58]
Table 5. Zebrafish models for MDs linked to membrane proteins.
Table 5. Zebrafish models for MDs linked to membrane proteins.
Gene SymbolProteinTechniqueRef.
itga7Integrin-α7MO[109]
cav3Caveolin-3Tol2 transposon system
MO; mRNA injection
[63]
[108]
dysfDysferlinMO[110,111]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tesoriero, C.; Greco, F.; Cannone, E.; Ghirotto, F.; Facchinello, N.; Schiavone, M.; Vettori, A. Modeling Human Muscular Dystrophies in Zebrafish: Mutant Lines, Transgenic Fluorescent Biosensors, and Phenotyping Assays. Int. J. Mol. Sci. 2023, 24, 8314. https://doi.org/10.3390/ijms24098314

AMA Style

Tesoriero C, Greco F, Cannone E, Ghirotto F, Facchinello N, Schiavone M, Vettori A. Modeling Human Muscular Dystrophies in Zebrafish: Mutant Lines, Transgenic Fluorescent Biosensors, and Phenotyping Assays. International Journal of Molecular Sciences. 2023; 24(9):8314. https://doi.org/10.3390/ijms24098314

Chicago/Turabian Style

Tesoriero, Chiara, Francesca Greco, Elena Cannone, Francesco Ghirotto, Nicola Facchinello, Marco Schiavone, and Andrea Vettori. 2023. "Modeling Human Muscular Dystrophies in Zebrafish: Mutant Lines, Transgenic Fluorescent Biosensors, and Phenotyping Assays" International Journal of Molecular Sciences 24, no. 9: 8314. https://doi.org/10.3390/ijms24098314

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop