Next Article in Journal
Expression and Transcript Localization of star, sf-1, and dax-1 in the Early Brain of the Orange-Spotted Grouper Epinephelus coioides
Next Article in Special Issue
Drosophila as a Model System for Studying of the Evolution and Functional Specialization of the Y Chromosome
Previous Article in Journal
Better Agreement of Human Transcriptomic and Proteomic Cancer Expression Data at the Molecular Pathway Activation Level
Previous Article in Special Issue
Drosophila Trachea as a Novel Model of COPD
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Inter-Species Rescue of Mutant Phenotype—The Standard for Genetic Analysis of Human Genetic Disorders in Drosophila melanogaster Model

by
Alexandru Al. Ecovoiu
1,
Attila Cristian Ratiu
1,*,
Miruna Mihaela Micheu
2 and
Mariana Carmen Chifiriuc
3
1
Department of Genetics, Faculty of Biology, University of Bucharest, 060101 Bucharest, Romania
2
Department of Cardiology, Clinical Emergency Hospital of Bucharest, 014461 Bucharest, Romania
3
The Research Institute of the University of Bucharest and Faculty of Biology, University of Bucharest, 050095 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(5), 2613; https://doi.org/10.3390/ijms23052613
Submission received: 2 February 2022 / Revised: 23 February 2022 / Accepted: 24 February 2022 / Published: 27 February 2022

Abstract

:
Drosophila melanogaster (the fruit fly) is arguably a superstar of genetics, an astonishing versatile experimental model which fueled no less than six Nobel prizes in medicine. Nowadays, an evolving research endeavor is to simulate and investigate human genetic diseases in the powerful D. melanogaster platform. Such a translational experimental strategy is expected to allow scientists not only to understand the molecular mechanisms of the respective disorders but also to alleviate or even cure them. In this regard, functional gene orthology should be initially confirmed in vivo by transferring human or vertebrate orthologous transgenes in specific mutant backgrounds of D. melanogaster. If such a transgene rescues, at least partially, the mutant phenotype, then it qualifies as a strong candidate for modeling the respective genetic disorder in the fruit fly. Herein, we review various examples of inter-species rescue of relevant mutant phenotypes of the fruit fly and discuss how these results recommend several human genes as candidates to study and validate genetic variants associated with human diseases. We also consider that a wider implementation of this evolutionist exploratory approach as a standard for the medicine of genetic disorders would allow this particular field of human health to advance at a faster pace.

1. Introduction

Advances in animal model-based research markedly increased our understanding of molecular mechanisms that regulate physiological and pathological processes. Perhaps one of the greatest achievements in this field is the development of genetically engineered animal models which offer a valuable platform for disease modelling and testing of potential therapeutic strategies. Accordingly, choosing reliable animal models represents a critical step to speed up the successful integration of precision medicine into daily clinical practice [1]. With its short generation time, low cost, large brood size and ease of genetic manipulation, Drosophila melanogaster (the fruit fly) has emerged as a key organism to explore disease-related genetic mechanisms [2].
Homo sapiens and D. melanogaster share strong similarities regarding many biological functions such as reproduction, embryo development, locomotion, respiration, circulatory system and neurodevelopment [3,4,5]. This relies on a high degree of evolutionary conservation of important genomic features such as genes, core regulatory mechanisms and genetic pathways.
These analogies endorse phenotypic rescue experiments conceived to reveal inter-species functional gene orthology. A crucial rescue assay for modelling a human genetic disorder (hGD) in D. melanogaster is the functional complementation (heterologous rescue) of an appropriate mutant fruit fly strain by the orthologous human or mammalian transgene associated with the respective hGD. If the two genes or proteins with similar nucleotide or amino acid sequences are also functionally related, namely, if the molecular functions are evolutionary conserved, the human wild-type allele (hWT) of the gene of interest (hGOI) is expected to rescue, at least partially, the fruit fly mutant phenotype. In other words, to rescue or save a phenotype means to restore it to wild type with a transgenic copy of the orthologous gene from other species.
In the technical jargon, various synonyms are used for the inter-species rescue of the mutant phenotype. For example, scientists working with a humanized yeast model (Saccharomyces cerevisiae) coined the terms cross-species functional complementation, standing for heterologous rescue, and heterologous expression, meaning the ectopic expression of a human gene in yeast strains, regardless of whether the yeast is wild type or mutant [6,7]. Herein, we conversely utilize the terms heterologous rescue and functional complementation used by FlyBase [8], but the more intuitive phrase phenotype rescue (such as lethality rescue) is also used whenever appropriate. The rescue term is also used for saving a fruit fly mutant phenotype with a dWT (Drosophila wild-type) transgene to confirm that the phenotype is indeed determined by the presumed gene and not by a hidden mutation present in the genetic background. We will further refer the phenotype rescue with dWT as intra-specific rescue, to differentiate from inter-specific, heterologous rescue or functional complementation equivalent terms. A different type of rescue is the chemical rescue, which is not a genetic one but instead is part of an endeavour to find chemicals able to alleviate or save mutant phenotypes, allowing the screening for potential new drugs.
D. melanogaster is suitable for different modelling approaches of human genetic diseases. One strategy involves targeted mutagenesis of Drosophila gene of interest (dGOI) in conserved sequences shared with hGOI, or RNAi inactivation of dGOI, to reproduce phenotypes resembling pathologic aspects of the hGD in D. melanogaster. An alternative is the replacement of dGOI with a disease-specific allele of hGOI to mirror clinical phenotypes in D. melanogaster. Last but not least, a different avenue is to introduce into fruit flies either a wild-type or a mutant copy of a hGOI having no evident structural ortholog in the D. melanogaster genome, but which may be useful to reproduce in vivo some molecular interactions important for understanding of the hGD. Whichever experimental alternatives are to be considered in practice, either individually or overlapping, a key step is to perform preliminary inter-specific phenotype rescue experiments, namely, to check if the wild-type copy of hGOI is able to functionally compensate a mutant allele of the orthologous dGOI. A positive result shows evolutionary functional conservation between the two species and reinforces D. melanogaster as a suitable experimental platform for modelling that particular hGD. On the other hand, a failure of the inter-specific phenotype rescue attempt, either an intrinsic or a false negative one, may induce the geneticists to decide on a mammalian model.
Since the experimental strategies used to model hGDs on D. melanogaster are already detailed in a few excellent papers [9,10], we choose to focus on various examples of inter-species phenotype rescues relevant for the medical research.
High-quality sequenced and assembled genomes have become increasingly available and allow experts to identify genes and regulatory sequences relevant for human medical research. This achievement relies on a deep comparative scanning of the two genomes with state-of-the-art bioinformatics tools. If structural orthologous gene pairs of interest are identified, targeted mutagenesis may be induced in D. melanogaster by an array of highly effective methods developed for this experimental model. The mutant alleles are then subjected to genetic analysis methods to check for functional orthology to human genes responsible for the aberrant phenotypes underlying the respective hGD.
Briefly, the modus operandi of this experimental approach is the identification of a human gene associated with the hGD of interest, bioinformatics comparative analysis to scan D. melanogaster genome for a candidate structural ortholog of the human gene, generation and analysis of relevant mutant alleles in D. melanogaster, delivery of the orthologous human cDNA into the appropriate fruit fly mutant background by means of effective molecular constructs and checking for partial or complete rescue phenotype of the transgenic fruit flies. Commonly, the heterologous rescue experiments target mutant phenotypes determined by loss-of-function (LOF) alleles, which are either hypomorphic alleles with reduced activity or null alleles with no residual activity [11]. The experimental steps of heterologous rescue, which rely on the modular and versatile UAS-GAL4 system, are outlined in Figure 1 [12,13].
As an example, if the LOF allele is a recessive null lethal one, very young heterozygous mutant embryos are microinjected with an insertional vector containing the orthologous hGOI cloned under an UAS enhancer control. Transgenic adults containing both the LOF allele and UAS–hGOI construct are crossed with a strain containing both the LOF allele and a GAL4-driver with either specific or generic pattern of expression. If, in the F1 generation of this cross, the LOF/LOF homozygous individuals are viable, the heterologous rescue was successful due to activation of UAS-hGOI by GAL4. Therefore, the structural ortholog’s genes are also functionally orthologous, indicating that at least some of their functions were conserved during evolution.
A successful heterologous rescue result is a very strong indicator for functional orthology between the members of human–fruit fly gene pairs. It is important to mention that any functional improvement of the mutant phenotypes of LOF flies such as rescue of lethality, proceeding through a later developmental stage, increased lifespan, increased fertility, improved behavior, etc., deserves attention and qualifies the functionally rescued dGOI as an attractive candidate for modeling hGDs in D. melanogaster [14]. Even when heterologous rescue of mutant fruit flies was not performed with a hWT but with a mammalian orthologous gene [15,16], this functional conservation is a strong genetic logic to start a research project on that gene model [14,17].
FlyBase reports the rescue experiments as “heterologous rescue” in the Overview tab of the report of a human genetic disorder modelled in D. melanogaster. The link to the respective fruit fly orthologous gene opens a Gene Report webpage which contains a Functional Complementation Data Table. If functional data are available, links under the Ortholog tabs showing functional complementation and Supporting References are present. In practice, many studies reported in FlyBase describe rescuing of abnormal phenotypes induced by RNAi suppression of GOI, but care should be taken when interpreting such data. A recent report dealing with the difficulties arising from the RNAi method reveals that residual functional activity of some genes in D. melanogaster still exists even when this technology is improved [18].
We reviewed data obtained from heterologous rescue experiments supporting human–fruit fly functional gene equivalence and their value for genetic analysis of hGDs. To this end, we present relevant examples of neurodegenerative and neuromuscular disorders, cardiac pathophysiology, cancer and infectious diseases. To our best knowledge, the present paper is the first attempt to scrutinize up-to-date scientific literature and FlyBase (FB2021_06) for the vast majority of the heterologous rescue experiments performed in D. melanogaster. We argue that preliminary experiments of mutant phenotype rescue should be the paradigm for any relevant genetic analysis of hGDs on the D. melanogaster model.

2. Neurodegenerative and Neuromuscular Disorders

For more than 20 years, D. melanogaster has been employed to tackle neurodegenerative and neuromuscular human afflictions [19,20]. Due to its relatively complex brain, which harbors around 300,000 neurons organized into specialized areas with discrete functions [21], D. melanogaster displays complex behaviors such as learning, memory, depression, anxiety, competitiveness, aggressiveness and alcoholism. Consequently, D. melanogaster represents a valuable system for the study of neuronal dysfunction and related disorders particular to several neurodegenerative diseases such as Alzheimer’s disease (AD), amyotrophic lateral sclerosis (ALS), Angelman’s syndrome (AS), autism spectrum disorder (ASD), Charcot–Marie–Tooth (CMT) disease, Friedrich’s ataxia (FA) and Parkinson’s disease (PD), to name a few [22].

2.1. Parkinson’s Disease

PD is one of the most common neurodegenerative diseases, accompanied by specific tremors and slow movement caused by degradation of dopaminergic (DA) neurons in the midbrain. To decipher the poorly understood mechanisms of selective degeneration of DA neurons, interactions between the products of human α-Synuclein (α-Syn), parkin RBR E3 ubiquitin protein ligase (PRKN) and PAELR genes were modelled in D. melanogaster brain neurons [23]. Co-expressing PRKN and α-Syn in transgenic flies rescued the loss of DA neurons and reduced the aggregation of α-Syn, a mutant phenotype, which endogenous parkin (park) from D. melanogaster was not able to rescue in α-Syn transgenic flies. This example can be viewed as a type of phenotypic rescue reflecting a putative incapacity of Drosophila’s park gene product to interact with human α-Syn.
On the same topic, Burchell et al. [24] showed that overexpression of human Fbxo7 gene, associated with a severe form of autosomal recessive early-onset PD [25], significantly rescued several mutant phenotypes such as locomotor defects, DA neuron loss and muscle degeneration determined by LOF of parkin. Pathogenic mutant Fbxo7 alleles were not able to rescue the loss of park, a result reinforcing the notion that the corresponding proteins share a common role in mitochondrial maintenance and mitophagy.
Mutations in coiled-coil-helix-coiled-coil-helix domain containing 2 (CHCHD2) human gene negatively impact the oxidative phosphorylation processes in mammalian cells [26] and are associated with an autosomal dominant form of late-onset PD. In D. melanogaster, LOF and hypomorphic Chchd2 alleles affect the maintenance of mitochondrial crista structure and lead to neuronal phenotypes associated with PD, such as sensitivity to oxidative stress, motor dysfunction, short lifespan and loss of DA neurons with age [27]. Expression of either transgenic dWT Chchd2 or hWT CHCHD2, but not missense alleles of the latter, successfully rescued the mitochondrial morphology and DA neurons loss induced by hypomorphic Chchd2H43 in D. melanogaster.
Another study concerning PD [28] focuses on exploring a specific subset of human iPLA2-VIA/PLA2G6 mutations that direct α-Syn aggregation and DA neurodegeneration specific for the PARK14-linked PD with α-synucleinopathy. The iPLA2-VIA/PLA2G6 gene codifies for an enzyme that is fundamental to phospholipids synthesis by the remodeling pathway or Lands’ cycle [28]. IPLA2-VIA-null allele impacts the early developmental stages of Drosophila mutants and leads to alterations of neurotransmission and midbrain DA neurons’ degeneration, causing gradual locomotor defects and sleep disruption [28]. When transgenic hWT iPLA2-VIA was expressed in the neurons of mutant flies, the motor and paralytic phenotypes were rescued, pointing to functional conservation between the two orthologous genes.
In addition to the previous examples, complex molecular interactions characterizing PD, such as the imbalance in trace metal levels characteristic for some forms of PD and AD, may be also addressed. The metal-responsive transcription factor 1 (MTF-1) gene is evolutionary conserved between D. melanogaster and mammals [29] and counteracts the effects of heavy metal loads. In mammals, MTF-1 was found to induce transcription of specific target genes in response to oxidative stress and infection [30]. A study focusing on the interactions between metal homeostasis and park function established that mutants expressing both park and MTF-1 LOF alleles in homozygous condition define a genetic assembly termed synthetic lethality [31]. The introduction of a transgene of MTF-1 in the double homozygous park and MTF-1 mutants rescued the lethality and has significantly increased the lifespan of park homozygous mutants. Alternatively, human MTF-1 has been able to rescue the short lifespan phenotype of park mutants [31], and largely, but not completely, rescued the metal sensitivity characterizing the LOF MTF-1 flies [32].
Loss of function pink1 mutant flies experience PINK1 deficiency and display motor disturbances as well as corrupted function of Complex I of the mitochondrial respiratory chain, and thus increased sensitivity to apoptotic stress. In humans, mutations in PTEN (phosphatase and tensin homologue)-induced kinase1 (PINK1) are strongly correlated with recessive forms of PD. HWT allele, but not mutant PINK1, rescued the phenotype exhibited by LOF Pink1 mutant flies [33].

2.2. Amyotrophic Lateral Sclerosis

ALS is arguably the most prevalent motoneuron disorder that leads to fatal adult-onset neurodegenerative progression [34,35]. The genetic basis of ALS overlaps at least partially with that of frontotemporal dementia, and often, the ALS patients concurrently develop cognitive and behavioral alterations [36,37].
Among over 30 genes that could harbor ALS causing mutations, some of the most noticeable genes are superoxide dismutase 1 (SOD1), chromosome 9 open reading frame 72 (C9orf72), fused in sarcoma RNA binding protein (FUS) and TAR DNA-binding protein (TARDBP) [38].
Various studies focused on validating the effects of human SOD1, as well as of other ALS associated genes, were performed on the fly phenotype in order to establish an ALS experimental model, as reviewed elsewhere [38,39,40]. In D. melanogaster it has been shown that null alleles of resident Sod1 determine impaired locomotor activity and lethality. These severe phenotypes are fully rescued by SOD1WT but not by its clinically relevant mutant alleles such as SOD1A4V, SOD1G37R or SOD1I113T [41], which confirm the impaired functions of these alleles in human patients. In addition, it was revealed that even a localized adult motor neuron expression of SOD1WT restored the lifespan of null Sod1-null flies to 60% of the normal controls [42].
In humans, both TARDBP and FUS code for DNA- and RNA-binding proteins involved in RNA processing of thousands of targets and share some common functionality underlined by similar pathogenic outcomes stemming from mutations [43,44,45]. D. melanogaster harbors TAR DNA-binding protein-43 homolog (TBPH) and cabeza (caz) as orthologs of TARDBP and FUS, respectively. LOF alleles of TBPH cause the disruption of mitochondrial trafficking accompanied with severe motor dysfunctions revealed by low rates of eclosion, altered larval crawling and adult climbing capacity [46]. Except the adult climbing mutant phenotype, all of the others are fully rescued by expressing either TBPHWT or TARDBPWT transgenes. Intriguingly, the mitochondrial transport defects were also rescued by expressing TARDBPM337V, an ALS-linked allele. This particular example of phenotypic rescue reveals that the pathogenic variant can display normal function in D. melanogaster, indicating that this allele is not involved in ASL. The same study found that caz1-null mutants [47] presented a significant decrease in mitochondria and vesicle transport. These phenotypes were rescued by expressing FUSWT in mutant flies, but the cazWT transgene was able to fully rescue both phenotypes only at 29 °C, when it is overexpressed, and only partially at 25 °C. Surprisingly, the FUSP525L pathogenic allele as well as its equivalent cazP938L successfully rescued the mitochondrial transport defects but not the vesicle transport. Regardless of expressing WT or pathogenic FUS alleles, other phenotypes particular to caz1 mutants such as eclosion, larval crawling and climbing defects were fully rescued. To cement the functional overlap between TBPH and caz, the reduced viability, lifespan, eclosion and climbing ability of TBPH mutants were fully rescued by neuronal overexpression of caz, but not vice versa. Overexpression of cazWT did not rescue the mitochondrial transport defects or larval crawling impairment. Consistent with these findings, it was previously shown that TARDBPWT transgene is also able to rescue the drastically reduced locomotor speed of mutant flies lacking TBPH, just as overexpressing cazWT [47].

2.3. Autism Spectrum Disorder

ASD comprises of complex developmental conditions and it is mainly characterized by behavioral symptoms such as impaired communication skills, defective social interaction, repetitive behavior, limited capacity to live independently, etc. but also by a high prevalence of gastrointestinal problems such as diarrhea, constipation, vomiting, and abdominal pain, just to mention a few [48,49,50,51,52].
SFARI Gene (https://gene.sfari.org/, accessed on 20 December 2021) is a database indexing the genes associated with ASD and categorizes these genes according to evidence of their involvement in ASD [53]. Within Category 1 of high confidence for implication in ASD, there are currently 207 genes, which are also present in other similar gene lists or were previously identified in an extensive exome sequencing study [54]. Out of these, 203 have orthologs in D. melanogaster, with 141 genes having a Drosophila RNAi Screening Center integrative ortholog prediction tool (DIOPT) score of at least of 0.6 [55]. DIOPT scores are provided by the DIOPT integrative tool [56], which is currently at version 8.5 and enables the search of orthologs in different species among the data provided by 18 large-scale ortholog prediction tools. The aforementioned list of orthologs with conclusive DIOPT scores includes the D. melanogaster genes Fmr1, Pten or ubiquitin protein ligase E3A (Ube3a), with DIOPT scores of 0.73, 0.87 and 0.93, respectively [55].
In D. melanogaster, Fmr1 gene is a structural ortholog of FMRP translational regulator 1 (FMR1) human gene. Mutations in FMR1 gene are causing human fragile X syndrome (FXS), which is probably the most common heritable foundation of autism disorders and mental retardation [57]. The two paralogs of FMR1, FMR1 Autosomal Homolog 1 (FXR1) and FXR2, also share a strong sequence similarity with Fmr1, thus making it difficult to choose a certain gene for testing the functional orthology. The Fmr150M-null allele determines a wide range of mutant phenotypes, the most striking in neurons and germ cells of adult flies. All three paralog human genes were tested in order to assess whether mutant phenotypes can be rescued in Fmr1-null mutants. Targeted neuronal expression of both hWT FMR1 and dWT Fmr1 transgenes rescued the characteristic FXS phenotypes such as higher brain protein levels, abnormal circadian rhythm patterns determined by small ventrolateral neurons’ synaptic arbor overgrowth defect and increased synaptic branching of neuromuscular junction. Expression of FXR1 and FXR2 failed to rescue the neuronal mutant phenotypes, however, all three human paralogs were equally competent to overcome non-neuronal symptoms in Fmr1 mutants such as severely reduced fertility caused by immotile sperm exhibiting defects in sperm tail microtubule organization [58]. These results highlight that care should be taken when testing structurally similar functional orthology candidates, especially when working with genes that manifest both evolutionary conserved and shared roles.
The PTEN gene has tumor suppressor activity, and even a partial loss of PTEN activity leads to cancers [59] or PTEN hamartoma syndrome, consisting of a variety of disorders such as macrocephaly, epilepsy, mental retardation and ASD [60,61]. In mammals as well as in D. melanogaster, PTEN, which is a dual lipid and protein phosphatase, is a critical repressor of phosphoinositide 3-kinase (PI3K)/protein kinase B (PKB/AKT) pathway [62]. In D. melanogaster, Pten100/Pten117 represents a strong hypomorphic heteroallelic combination that leads to increased larval growth and, consequently, increased pupal volume and adult weights. Expression of the PTENWT allele in embryos and larvae successfully rescued the hypomorphic mutants [63]. Within the same study, the authors devised a scalable experimental platform to functionally test about 100 human PTEN alleles with potential clinical relevance. To do this, they overexpressed an activated Phosphatidylinositol 3-kinase 92E (Pi3K92E) allele (PI3K92E–CAAX–PI3Kact) in the wing imaginal disc, resulting in adults harboring enlarged wings. This phenotype was rescued by PTENWT but not by PTENC124S, which lacks both protein and lipid phosphatase activity, or by the PTENG129E lipid phosphatase dead alleles. PTENY138L, which is deficient for protein phosphatase activity, is able to partially alleviate the wing size phenotype. These either failed or partially successful heterologous rescue experiments demonstrated that PTEN-dependent suppression of PI3K/AKT tissue growth in Drosophila is dependent upon both lipid and protein phosphatase activities. Using this wing size-based system, they successfully tested the functionality retained by the PTEN alleles by scoring their ability to partially or completely rescue the oversized fly wing.
Mutations affecting UBE3A gene are the main cause for AS, a relatively common human disorder involving aberrant central nervous system development and characterized by mental retardation and locomotor impairment [64]. In our laboratory, we obtained a new Ube3a allele, symbolized as Asm1.5-R, consecutive to transposon-mediated mutagenesis [65]. This allele is semilethal for mutant homozygous males, which, when surviving to adulthood, elicit decreased locomotor performances. We successfully rescued this abnormal phenotype in flies raised for more than a year on culture medium supplemented with omega-3 polyunsaturated fatty acids, i.e., eicosapentaenoic and docosahexaenoic acids [66]. A true heterologous rescue was demonstrated for the learning abilities of both larvae and adult Ube3a-null mutants [67]. Individuals from both mutant categories exhibited impaired learning abilities as scored by using the aversive phototaxis suppression assay, which tests the ability of fruit flies to link an unpleasant taste stimulus with light. Expressing UBE3AWT transgene by the pan-neuronal elav-GAL4 driver in the mutant background rescued the mutant learning defects.
Altogether, the previous rescue and disease modelling examples reinforce the power of the D. melanogaster experimental model. In addition, many other mammalian genes associated with neurodegeneration-related diseases were studied in the fly model, as presented in Table 1.

3. Cardiac Disorders

Several key characteristics related to cellular processes, signalling pathways and gene conservation endorse D. melanogaster as a model of choice for studying human cardiac development, function and diseases. First, although seemingly simplistic, the fruit fly circulatory system, consisting of a tube-like heart that pumps the haemolymph, shows developmental and functional similarities to the vertebrate heart [134,135,136,137]. Second, both organisms (D. melanogaster and H. sapiens) share some regulatory cardiogenic networks encompassing critical cardiac transcription factors such as tin/Nkx2.5, Mef2/Mef2C, pannier/GATA family and Hand/HAND1 and HAND2, which are required for cardiac progenitor specification [137,138,139,140,141,142,143]. Third, there is a strong gene conservation with human genes, particularly with disease-related genes [144]. By performing a systematic BLAST analysis of 929 human disease gene entries associated with at least one variant in the Online Mendelian Inheritance in Man (https://www.omim.org/, accessed on 18 December 2021) database against the reference sequence of D. melanogaster, Reiter and colleagues revealed that 77% of disease genes queried had Drosophila counterparts [144]. Of note, 26 of them were associated with various cardiovascular diseases such as cardiomyopathies, hypertension and conduction defects [3].

3.1. Congenital Heart Defects

Congenital heart defects (CHDs) are the most common birth disorders, affecting 0.8% to 1.2% of live infants [145]. Although it is largely acknowledged that genetic factors are strongly involved in CHD pathogenesis, the great majority of responsible genes remain elusive [146]. Drosophila-based research enabled the recognition of new CHD-related genes. Zhu et al. [147] developed a Drosophila-based functional system to rapidly and efficiently screen large numbers of candidate genes detected in patients with severe CHDs. By using heart-specific RNAi silencing, they tested 134 genes, of which more than 70, including a subgroup encoding histone modifying proteins, were found to be essential for the development, structure and function of the fruit fly heart. The silencing of genes responsible for H3K4 and H3K27 methylation (i.e., kis/CHD7, wds/WDR5, Trx/MLL2) caused developmental lethality (up to 84%) and severe structural heart anomalies and reduced adult longevity. Moreover, a gene substitution strategy comprising concurrent heart-specific silencing of the fly gene homolog and expression of either a wild-type variant or a pathogenic one was applied to validate the role of these genes in CHDs. As a proof-of-concept, the authors explored the potential of the WDR5WT human allele to rescue the pathogenic phenotype generated by silencing of the endogenous wds Drosophila homolog. WDR5WT overexpression significantly reduced developmental lethality and restored abnormal heart morphology, as opposed to the CHD patient-derived WDR5K7Q mutant allele, which resulted in similar pathogenic cardiac manifestations.

3.2. Cardiomyopathy Phenotypes

Cardiomyopathies are a heterogeneous group of myocardial diseases in which the cardiac muscle is structurally and functionally abnormal, often due to a genetic cause [148]. The injury can be limited to the heart or part of a generalized systemic disorder; either way, the genetic architecture is very diverse [149,150].
Hypertrophic cardiomyopathy (HCM) is the most prevalent inherited cardiomyopathy, affecting at least 1 in 500 individuals in the general population [151,152]. The underlying genetic etiology is complex, mainly involving variation in sarcomeric or sarcomeric-related genes, but mutation in other genes can cause similar phenotypes comprising left ventricular hypertrophy (LVH). Of the 57 candidate genes included in diagnostic HCM gene panels, only 8 have been nominated as having definitive evidence, including myosin light chain 2 (MYL2) [153,154].
Recently, Manivannan and colleagues identified a novel recessive frameshift variant in MYL2 (p.Pro144Argfs*57) resulting in early-onset HCM and death in infancy [155]. A fly model was used to demonstrate that p.Pro144Argfs*57 variant was in fact a LOF allele. The expression of the Drosophila ortholog Mlc2 was knocked down using transgenic RNAi lines, which led to multiphasic lethality, with progenies dying before the pupal stage, and impaired systolic function. Both the developmental lethality and cardiac dysfunction were partially rescued by MYL2WT but not by the frameshift variant. The incomplete restoration of fly phenotype was most likely due to sequence differences between the two organisms, given that the Mlc2 N-terminal region has additional sequences that are required for its function [156].
An HCM-like phenotype can be encountered in other conditions involving LVH, such as FA, which is a neurodegenerative disease caused by a GAA trinucleotide repeat expansion in frataxin gene (FXN) [157,158]. Over time, the cardiomyopathy potentially progresses to a dilated form. It has been shown in D. melanogaster that RNAi-mediated frataxin (fh) depletion prompted enlargement of cardiac diameters and reduction in systolic function, which were fully rescued by complementation with FXNWT human allele [159].
Gonçalves et al. [160] reported a study concerning three human families A, B and C affected by mutations in adducin 3 (ADD3) gene encoding for adducin-γ, associated with various disabilities such as intellectual disability, microcephaly, cataracts and skeletal defects. Patients from family A are also homozygous for a missense mutation of lysine acetyltransferase 2B (KAT2B) and the expanded pathologic spectrum, including cardiomyopathy and renal problems. This allele is symbolized as KAT2BF307S and determines the substitution of highly conserved Phe with Ser at position 307 of the human protein. These genetic disorders are prone to be modeled on D. melanogaster, since the fruit fly genome contains the hu li tai shao (hts) ortholog for ADD1 (adducin-α), ADD2 (adducin-β) and ADD3, and Gcn5 acetyltransferase (Gcn5) paralog for KAT2A and KAT2B.
The htsnull hemizygous flies die as late larvae and only a few impaired, short living escapers reach the adult stage. Expression of ADD3WT does not rescue viability, but the ubiquitous co-expression of ADD1WT and ADD3WT leads to heterologous rescue by increasing the number of viable adults. Alternatively, ADD1WT/ADD3E659Q co-expression (where ADD3E659Q is a human mutant allele reported for family A) induces only a partial rescue of htsnull hemizygous flies, revealing that ADD3E659Q behaves as a hypomorphic allele in the fruit fly mutant background. The hts mutant flies rescued by ADD1WT/ADD3E659Q did not show any significant differences in heart period, cardiac output, fractional shortening and arrhythmia index as compared to those rescued with ADD1WT/ADD3WT.
The genetic analysis of an allele of Gcn5 gene in D. melanogaster equivalent to KAT2BF307S supports the hypothesis that KAT2B is associated with heart and kidney mutant phenotypes in humans. Specifically, the Gcn5E333st allele, also referred as Gcn5null, is lethal in hemizygous individuals, which arrest development at the late larval to early pupal stage [161]. When KAT2AWT and KAT2BWT are expressed either individually or synchronously in hemizygous Gcn5null flies, the rescue of transgenic flies fails, suggesting that these orthologous genes have functionally diverged during evolution. As expected, transgenic hemizygous Gcn5null flies appear to be completely rescued by the Gcn5WT allele, but partially rescued by the Gcn5F304S allele, as lethality still often occurs in pupae or in early adults and the escapers exhibit morphological impairments. Gcn5F304S resembles KAT2BF307S and encodes a protein variant having Ser instead of Phe at position 304. Remarkably, the escapers expressing Gcn5F304S not only have obvious morphological mutant phenotypes, but also present a prolonged heart period and reduced cardiac output comparative to both a control strain and Gcn5null hemizygous flies rescued with Gcn5WT.
RNAi silencing of Gcn5 (Gcn5RNAi) in D. melanogaster induces functional heart problems, while silencing of hts (htsRNAi) does not. However, silencing of both genes in Gcn5RNAi and htsRNAi flies aggravates heart period length and the arrhythmia index induced by Gcn5 knockdown alone. These data reveal that Gcn5 is directly involved in heart function in D. melanogaster, while some hts mutations increase the severity of the Gcn5-null phenotype [160], spotting hts as a potential genetic enhancer of Gnc5.
Such complex heterologous rescue experiments of specific D. melanogaster mutants simulate patients with multilocus genetic diseases, affected by pathogenic mutations located in more than one gene. In this case, simultaneous knockdown of hts and Gcn5 concurrently increased the severity of heart phenotype in fruit flies, but the heterologous rescue succeeded only for hts. Nevertheless, when this partial success is corroborated with cardiac phenotypes reported for the equivalent KAT2BF307S and Gcn5F304S alleles and the RNAi results, the perspectives are encouraging. It is reasonable to conclude that various interactions between ADD3 and KAT2B variants in patients can be mirrored by interplays of hts and Gcn5 alleles in D. melanogaster, helping experts to develop novel drugs able to restore the normal cardiac phenotype.
Intriguingly, a closer inquiry revealed that Gcn5WT rescues lethality and the morphology of wings, legs and eyes of hemizygous Gcn5null flies, but, similar to Gcn5F304S transgenic individuals, these organisms have a smaller diastolic diameter as compared to control flies. A possible explanation for this phenotype is that Gcn5WT transgene is not in its natural genomic environment where its regulators reside. Nevertheless, when compared to the Gcn5WT rescued flies, the Gcn5F304S transgenic ones exhibit supplemental cardiac impairments as reduced contractility and a more irregular heartbeat. This case is a very interesting one, as it shows that subtle phenotypes may still be present even when a complete intraspecific rescue is reported. It seems that sporadic complete phenotype rescue results may remain partial to some degree, as subtle mutant phenotypes may be difficult to notice unless specifically searched for, as exemplified in the study of Gonçalves et al. [160].
Fundamental questions emerge when considering that KAT2AWT and KAT2BWT counterintuitively fail to rescue hemizygous Gcn5null flies, pointing to a functional divergence. Why are KAT2BF307S and Gcn5F304S equivalent alleles associated with similar cardiac phenotypes in humans and flies, suggesting an inter-specific functional conservation? Should one always expect that a structural gene orthology is concluded by heterologous rescue experiments? Why does functional complementation of hemizygous Gcn5null flies with human WT alleles fail? Considering that most heterologous rescue experiments reported for D. melanogaster were performed using the GAL4-UAS system, it is helpful to consider the recent work of Casas-Tintó et al. [162]. Due to carefully designed experiments, the authors conclude that the expression of enhancer-Gal4 constructs may be transiently ectopic and influenced by the genomic insertion site. Added to the fact that not a complete human gene sequence but a human cDNA, without regulatory sequences, is usually cloned in a UAS vector, we presume that the unstable activation history of some enhancer-Gal4 constructs may interfere with the heterologous rescue results.

3.3. Other Cardiac Disorders

Other cardiac disorders have been modelled in D. melanogaster, such as channelopathies [163,164,165,166,167] and different syndromic [168,169,170] or nonsyndromic cardiomyopathies [171,172,173,174,175,176,177]. To our knowledge, currently, none of these diseases benefit from effective human allele-based functional complementation studies, although some groups successfully tackled the heterologous rescuing of fly cardiac phenotype, the alleles of choice being of animal origin, mainly from mice [173,178,179]. For example, Gao’s group reversed the effect of loss of fly γ-sarcoglycan (Scgδ) by using a murine counterpart gamma-sarcoglycan (Sgcg) [179]. An engineered form of the Sgcg (termed Mini-Gamma) has been introduced into flies, and it efficiently rescued the cardiac phenotype of an amorphic allele of Scgδ. Mini-Gamma was generated by removing a portion of extracellular domain of Sgcg that contained a large frameshift deletion, which led to a premature stop codon. Exon skipping corrected the reading frame, expression of Mini-Gamma in the heart tube being sufficient to restore cardiac function to wild-type magnitudes.
A gene involved in cardiac dysfunction independently from canonical Wnt signaling is pygopus (pygo), which maintains normal heart physiology in aging D. melanogaster [180] and is involved in the differentiation of intra-cardiac valves [181] of the fruit flies. Knockdown pygo mutant allele underpins cardiac arrhythmias and decreased contractility with systolic dysfunction in fruit flies [182]. The cardiac impairments determined by knockdown pygo allele in D. melanogaster resemble increased incidence of atrial fibrillation in senior humans. Although Pygo1 and Pygo2 are not essential for heart function and development in mouse, they may be involved in preventing senescence phenotypes specific for aging hearts in mammals [180].
An experiment of interest would be the functional rescue of the cardiac phenotype of pygo knockdown fruit flies with the transgenic Pygo1WT allele, a plausible scenario, as the lethality of pygo130-null embryos was rescued by both PYGO1WT and hPYGO2WT human alleles [182]. Again, the experimental paradigm is that if heterologous rescue of a specific severe phenotype such as lethality is possible, then rescuing subtle mutant phenotypes determined by the same orthologs is reasonably plausible.
The use of high-throughput sequencing techniques and wide-ranging cardiac gene panels dramatically increased the detection of variants of uncertain significance (VUS) [183,184,185,186], whose definite classification requires additional studies including functional ones. The previously presented data, as well as data from Table 2, demonstrate the structural and functional homologies between fruit fly and human cardiac genes and advocate the use of D. melanogaster system as a prime candidate to study and validate genetic variants associated with cardiac disorders.

4. Cancer

Cancer is a multifactorial and multistep disease characterized by uncontrolled proliferation of tumor cells that escape the control of physiological growth sentinels, apoptosis defects and metabolic alterations. These signaling pathways are conserved in D. melanogaster, making the fruit fly an appropriate model organism to study cancer biology [192,193,194,195]. Important processes such as genomic instability, strategies to evade apoptosis, telomerase reactivation, tumor-promoting inflammation and evasion from the immune system, angiogenesis, anaerobic glycolysis, competitiveness of cancer stem cells, importance of tumor microenvironment, invasiveness and metastasis, cancer cachexia, drug screening and resistance have been extensively studied and modeled in D. melanogaster [196]. Several examples are provided in Table 3 and the following subchapters.
Epithelial cancer is the most studied type of cancer on the D. melanogaster model, as revealed by the high number of citations recorded in FlyBase. This is at least partially explained by the fact that the fruit fly larval imaginal discs, which are morphologically and biochemically comparable to mammalian epithelia, could be used to model different processes involved in the epithelial cancer onset and progression. The imaginal wing and eye discs have been successfully used to study tumor growth and invasion, investigate the function of cancer genes, analyze oncogenic cooperation and perform chemical screenings [205,206,207,208,209].

4.1. Validating Orthologs of Human Tumor Suppressors Using the Drosophila melanogaster Model

In D. melanogaster, three complexes are involved in regulating cell growth and differentiation: Crumbs/Stardust/PATJ/Bazooka, Par6/aPKC (atypical protein kinase-C) and Scrib/Dlg/Lgl (Scribble/Discs large/Lethal giant larvae) complexes [210]. The lgl was the first neoplastic tumor suppressor gene discovered in Drosophila, whose loss leads to an abnormal development (disruption of cell polarity and tissue architecture, uncontrolled proliferation and tumor growth) of the imaginal structures and the larval brain. The apical–basal polarity loss in epithelial cells often occurs in human epithelial cancer, facilitating invasion and metastasis, and therefore a more aggressive profile of the malignancy [211]. Following their transplantation into wild-type recipients, the lgl mutant imaginal tumorous tissues could migrate and metastasize in other regions of the fruit fly body, killing the host, thus resembling the human secondary cancers [212]. Other features shared with human metastasis are represented by the upregulation of type IV collagenase and NDP kinase in lgl-induced tumors [213,214]. Mammalian homologues of the lgl gene (HUGL-1/Llgl1 and HUGL-2) are highly conserved in humans, highlighting their role in cell growth and initiation of neoplastic lesion. From the two human homologues of the lgl gene, the HUGL-1 LOF has been reported in different types of human cancer (e.g., breast, melanomas, prostate, ovarian and lung cancers) and HUGL-1 rescued all the defects of the fly lgl mutant. For the rescue experiments, the null allele lethal(2)gl4 has been used. The flies homozygous for the mutant allele are headless pharate, with the eye imaginal disc structure completely lost in the third instar larval stage [215]. The insertion of the HUGL-1 cDNA in the homozygous mutants led to a partial development of rudimental eyes and larval structures comparable to wild type. The ubiquitous expression of HUGL-1 in lgl-null fruit flies assured the recovery of viable phenotypes (viable adults or completely developed pharate). Despite being completely sterile, they did not develop neoplasia during their lifespan and showed normal imaginal structures, compared to that of the wild-type adults. These results demonstrate that HUGL-1 can act as a tumor suppressor in D. melanogaster and thus represents the functional homologue of lgl [197].

4.2. Elucidating the Role of Tumor Microenvironment and Host-Neoplastic Cells Competition in Gut Adenoma Development

It is largely accepted that the tumor microenvironment plays an important role in the tumor’s progression, exhibiting either pro-growth or inhibitory effect on the proliferation and invasiveness of neoplastic cells [216]. Cancer cells use cell competition as a form of interaction within the tumor microenvironment [217,218]. Cell competition was first described as a quality control mechanism in Drosophila defined as the ability of wild-type cells to kill the mutant cells harboring reduced fitness and growth potential [219]. However, normal cells could also be killed by tumor mutant cells called supercompetitor cells, leading to the development of hyperplasia and adenomas in the adult Drosophila midgut [220]. Studies in the Drosophila model have demonstrated that cooperation between the tumor suppressor adenomatous polyposis coli (APC) gene and wingless (Wnt) is involved in cell competition. Two APC proteins, APC1 and APC2, with different domains and tissue distribution are shared by mammals and Drosophila. Fly APC1 and APC2 functions are partially redundant in regulating Wnt signaling and cytoskeletal reorganization [221].
As also reported for humans, the APC inactivation in Drosophila leads to very high levels of Wnt target gene expression in different tissues, including the intestine. Akin to the mammalian intestine, Drosophila adult midgut epithelium cells have a high turnover rate maintained by the intestinal stem cells (ISCs). Therefore, they have been used as a model to elucidate the role of different signaling pathways in adenoma formation and the role of different mutations in tumor development [222,223,224,225,226]. Loss of function of APC leads to abnormal proliferation of ISCs, followed by the loss of gut epithelial cell polarity, hyperplasia and epithelial overgrowth [227,228]. On the other side, the Wnt genes are expressed at very high levels in colorectal tumors harboring mutations in APC. This proves that upregulation of Wnt expression in human ISCs is associated with adenoma development [229]. Importantly, loss of APC1 leads to the activation of Wnt signaling in retinal photoreceptors, inducing their ectopic apoptosis. Thus, the APC1 mutant eye phenotype could be used to investigate the roles of Wnt signaling pathways in health and disease [230,231]. Using this Drosophila model, it has been shown that the Wnt pathway genes’ expression in Drosophila is regulated by transcription cofactors such as earthbound1 (ebd1) and erect wing (ewg) [200,232]. From these, ebd1, which harbors Centromere Protein B (CENPB) DNA binding domains, is essential for the Wnt-dependent control of ISC proliferation. Using the APC1 mutant eye phenotype, it has been demonstrated that ebd1 heterozygotes induce a partial suppression of APC1 mutant apoptosis, the rescue being nearly complete in ebd1 homozygotes. The human homologue of ebd1 (JRK/JH8) is overexpressed in several carcinomas including colon, breast and ovarian serous cystadenocarcinoma, and has been proven to be associated with elevated expression levels of Wnt target genes in human colorectal tumors. ebd1 and its human ortholog share structural and functional similarity and interact directly with Arm/β-catenin associated in a ternary complex with T cell factor (TCF) [233]. The human Jerky has rescued the flight muscle defects in ebd1 as well as in ebd1 and ebd2 double mutants proving its functional equivalence to Ebd1 and Ebd2 [200].

4.3. Demonstrating the Species-Dependent Pathways of Notch Hyperactivation

Coiled-coil and C2 domain-containing protein (CC2D) 1A and 1B are members of the Lgd protein family, conserved among metazoans, with many partially redundant functions such as centrosomal cleavage, molecular signaling, innate immunity response (by modulating TLR3, TLR4 and RLR pathways) and synapse maturation [234,235]. The D. melanogaster ortholog Lgd, a tumor suppressor gene, has been shown to be involved in Notch signaling. Notch is a conserved developmental signaling pathway involved in essential cellular processes such as differentiation, pattern formation, cell-cycle progression, morphogenesis, migration, apoptosis, T cells activation etc., which is dysregulated in many cancer types [236]. The Notch signaling occurs upon the direct contact between a signal-receiving and a signal-sending cell, mediated by the binding of Delta/Serrate/LAG-2 (DSL) ligand to the trans-membrane Notch receptor. Upon binding, the extracellular domain cleavage allows the release of active receptor Notch intracellular domain (NICD), which accumulates in the nucleus and regulates downstream genes in concert with other proteins [237,238]. Loss of function of lgd in fruit flies led to constitutive ligand-independent activation of Notch in epithelial cells and to tissue hyperplasia [239,240,241,242,243,244,245]. However, this effect was not confirmed in mouse epithelial gut, suggesting that lgd genes are not involved in the Notch pathway hyperactivation in mammals. In humans, NICD nuclear accumulation was associated with increased tumor cell growth and cell survival and treatment failure in different types of malignancies such as breast, lung and pancreatic cancer [246,247,248,249]. The functional homology between human CC2D1A and CC2D1B and the fly lgd was demonstrated in rescue experiments. Among the two orthologs, one copy of CC2D1B was sufficient to completely rescue the mutant [201]. Moreover, the expression of CC2D1A and CC2D1B is under the control of the endogenous promoter of lgd [244].

5. Infectious Diseases

The D. melanogaster model provides a valuable tool for studying the molecular mechanisms of different infectious diseases, the processes involved in anti-infectious immunity and for pharmacological screenings, as thoroughly detailed in the recent review of Harnish et al. [246]. This is at least partially because many infectious agents often modulate highly conserved innate immunity pathways such as the Nuclear Factor kappa B (NF-κB) and c-Jun N-terminal Kinase (JNK) signaling, phagocytosis and apoptosis, which are also present in the fruit fly. We will present below some examples illustrating that D. melanogaster represents an appropriate experimental model to recapitulate Koch–Evans postulates [247] regarding the reproduction of human phenotypes for some infectious diseases.

5.1. Molecular Mechanisms of Neuropathological Effects Caused by Zika Virus NS4A Protein

Zika virus is an emerging mosquito-borne flavivirus closely related to Dengue and West Nile viruses [248]. Zika infection is associated with severe neurological symptoms and sequelae, such as Guillain–Barre syndrome and congenital microcephaly [249,250]. These neurological complications are explained by the interaction of the nonstructural 4A (NS4A) protein of Zika virus with ankyrin repeat and LEM domain containing 2 (ANKLE2), encoded by a gene associated with autosomal recessive microcephaly in humans [116,251]. Indeed, the ANKLE2 heterozygous mutations in humans are associated with infants’ severe microcephaly and later cognitive, neurological, intellectual and developmental deficits [252]. Interestingly, Ankle2 is involved in brain development in flies, inhibiting the neuroblast division in the third instar larval brain; the hypomorphic mutants (Ankle2A) have been proved to be pupal lethal and exhibited small brain volume phenotype [116,253,254].
The D. melanogaster model was used to demonstrate the physical interaction between Zika NS4A and human ANKLE2 and its pathological consequences. The ectopic expression of Zika NS4A in the developing third instar larva brains of Drosophila provoked a reduction in brain lobe volume, induced apoptosis and reduced neuroblast proliferation. This mutant phenotype was rescued by wild-type ANKLE2 but not by a microcephaly-associated ANKLE2 variant (ANKLE2Q782X) [255]. The expression of NS4A in fruit flies heterozygous for a hypomorphic allele of Ankle2 caused a more significant microcephaly in comparison to the condition induced in wild-type fruit flies [255]. Moreover, Ankle2 regulates the function of genes that control cell polarity during asymmetric division of neuroblasts, including lethal (2) giant larvae (l(2)gl), atypical protein kinase C (aPKC), bazooka (also known as par-3) and par-6 and VRK1, for which human orthologs have been described. These genes are involved, both in flies and humans, in neural stem cell self-renewal and production of neurons, while being also related to developmental brain disorders [256]. From these, the human VRK1 pathogenic alleles are associated with motor and sensory axonal neuropathy and microcephaly [257]. Mutations in the fly homolog of VRK1, ballchen (ball), induced the loss of neuroblasts in third star Drosophila larval brain. It has been shown that NS4A, having the same location as Ankle2 in the endoplasmic reticulum and nuclear envelope, could interact with ball (VRK1) to regulate brain size in flies. The NS4A expression mimicked the influence of the Ankle2-Ball (VRK1) pathway on the aPKC and l(2)gl proteins, which are critical for brain development. The microcephaly induced by NS4A expression has been rescued by removing a single copy of ball or l(2)gl, demonstrating that NS4A hijacks the Ankle2-ball (VRK1) pathway, affecting neuroblast division and brain development, leading to microcephaly [258].

5.2. Elucidating the Molecular Players in the Cytotoxicity of Cholera Toxin

Cholera is an acute diarrheal infection caused by ingestion of food or water contaminated with the spiraled bacterium Vibrio cholerae, causing 1.3–4.0 million cases and 21–143 thousand deaths each year [259]. The most important virulence factor of V. cholerae is cholera toxin, a typical AB toxin with ADP-ribosylating action. Cholera toxin stimulates cAMP production in the gut epithelial cells, generating the hypersecretion of water and electrolytes responsible of the specific clinical symptoms such as aqueous diarrhea and rapid, severe dehydration [260].
Drosophila model was used to investigate the molecular mechanisms of cholera toxin active (A) subunit. For this purpose, the gene for cholera toxin A subunit (CtxA) was expressed in the developing fly wing, causing a CtxA-dependent weight-loss phenotype. The mutant phenotype was fully rescued upon co-expression of an active form of Notch or wild-type Rab11, the most well-represented members of the Ras superfamily GTPases, involved in intracellular vesicle trafficking and signaling [261], and was significantly worsened when a dominant-negative form of Rab11 or Gαs was co-expressed. The CtxA expressed in the midgut epithelial cells affected the intestinal epithelial permeability, as revealed by the occurrence of gradual wasting and the smurfing phenotype after flies feeding with food dyed with FD&C blue dye#1 [262]. These phenotypes have been also rescued by co-expression of Rab11, suggesting that cholera toxin A subunit inhibits Rab11-mediated vesicle trafficking.

5.3. Molecular Mechanisms of Apoptosis Induced by Helicobacter pylori Cytotoxin-Associated Gene A

Helicobacter pylori infection affects about 50% of the human population and around 7% will develop gastroduodenal disease [263]. H. pylori is associated with various gastric pathologies, ranging from peptic ulcer to gastric adenocarcinoma and lymphoma [264]. The symptomatic infections are produced by strains harboring the cytotoxin-associated gene A (cagA), one of the main virulence factors [265]. Once delivered in the host cells, CagA is activated through phosphorylation by Src-family kinases and binds to SHP-2, a protein phosphatase encoded by the PTPN11 gene in humans. SHP-2 further activates signaling pathways downstream of receptor tyrosine kinases (RTKs), a class of receptors with a pivotal role in cancer invasion and metastasis [266,267]. By interfering with epithelial cell adhesion, polarity, migration and differentiation, CagA triggers malignant transformation of gastric epithelial cells, its oncogenic potential being confirmed in transgenic animals [268,269].
The ectopic expression of CagA in the epithelial cells of the D. melanogaster developing wing has been shown to induce a dose-dependent apoptotic effect, leading to significantly lower size wings. This phenotype was similar to that produced by the localized activation of the JNK pathway within the wing cells. The cagA-induced apoptosis was suppressed by co-overexpressing a dominant-negative form of Basket (Bsk), a Jun amino-terminal kinase (JNK) homolog, and enhanced by co-overexpressing of wild-type Bsk [270], suggesting that CagA is an important mediator of the activation of JNK signaling pathway during H. pylori infection [271,272].
Since JNK signaling is dependent on the Ras oncogene, it has been further investigated whether CagA can genetically interact with the constitutively active oncogenic variant of Ras called p.G12V or RasV12 [272,273]. If the expression of RasV12 alone in the fly eye has been shown to induce the formation of non-invasive tumors, its co-expression with CagA causes invasive tumors. In addition, CagA has been shown to interact with dlg1 and l(2)gl, which are considered neoplastic tumor suppressor genes [274].

5.4. Discovering Novel Candidates for Assessing Genetic Susceptibility to Different Infections

The malvolio (mvl) gene of D. melanogaster encodes a protein sharing a high homolog to natural resistance-associated human macrophage proteins (Nramps), which are integrated in the phagolysosomal membrane and function as cation transporters [275,276,277]. In D. melanogaster, mvl is expressed in macrophages and in differentiated neurons. The loss-of-function mutations lead to taste behavior defects caused by a reduction in the sensitivity of the gustatory circuits to stimuli. The ubiquitous expression of human Nramp-1 protein in mutant fruit flies can fully rescue the taste defect. Moreover, the taste behavioral defects can be suppressed when the fruit flies are grown on media supplemented with Fe2+ or Mn2+ for a minimum of 2 h before testing, sustaining the role of mvl in the transport of bivalent cations [278].
The polymorphisms of the Nramp-1 gene have been linked to susceptibility to tuberculosis and leprosy in human populations; therefore, the fact that human Nramp-1 can fully complement the defect in mvl makes D. melanogaster an attractive in vivo model system for Nramp-1 functions in different human infections [279,280].

5.5. Demonstrating the Functional Homology of Human Vasodilator-Stimulated Phosphoprotein (VASP) and Drosophila enabled

Drosophila enabled (ena), a dominant genetic suppressor of mutations in the Abelson (Abl) tyrosine kinase, is a member of Ena/human vasodilator-stimulated phosphoprotein (VASP) protein family, which is associated with actin filaments and focal adhesions. Ena is a specific substrate for Abl and also interacts with the SH3 domain of Abl [281]. Among its physiological roles, VASP interacts with Listeria monocytogenes Act A protein, which is required for the internalization of this facultative intracellular pathogen, mediating the reorganization of actin filaments. The Ena/VASP domain 1 (EVH1) and EVH2 share 58% and 31% homology between Drosophila ena and human VASP, respectively. Moreover, EVH1 is similar to the WP1 domain found in Wiskott–Aldrich syndrome protein, which is associated with cytoskeletal defects in T cells and platelets [282].
Using Drosophila model, it has been demonstrated that VASP rescues the lethal phenotype associated with ena LOF. The lethal ena mutant alleles which affected EVH1 and EVH2 and their capability to bind the focal adhesion protein zyxin and the Abelson kinase were characterized by cytoskeletal defects. A comparison between VASP and three ena transgenic lines, tested for their ability to rescue the ena null lethal mutants, showed that VASP partially rescued ena mutant lethality, 25–85% survival rate and normal development four weeks after eclosion as compared to 79–100% in case of ena transgene. Thus, VASP represents a functional substitute of ena, the two proteins also sharing the same subcellular distribution and the same expression pattern in mammalian cells. They are detected at the level of actin filaments and focal adhesion contact where they bind to zyxin and the SH3 domain of Abl. Finally, lethal Ena mutations have been identified in the most conserved domains of the Ena/VASP family. Further research of this protein family might provide new insights into the regulation of cytoskeleton changes during physiological and infectious processes [283]. The interaction of Listeria monocytogenes ActA with VASP may explain the activation of actin reorganization and bacterial cell internalization.

6. Discussion

We emphasize that an a priori validation of the functional orthology of an hGOI/dGOI pair by heterologous rescue should be the experimental paradigm for in vivo modeling of an HGD. This preliminary approach, even when partially successful, validates the evolutionary conservation of the matching genes and qualifies the respective HGD as an appropriate candidate to be modeled on the D. melanogaster platform [13].
Various experimental assays pointing to functional orthology sometimes offer rather indirect data. Although conceptually overlapped, modeling of a human genetic disease by expressing human variants and equivalent fruit fly alleles in D. melanogaster does not equate with heterologous rescue experiments. Sometimes, modelling data are not even supported by functional complementation data, as in the case of the KAT2B/Gcn5 gene pair, described above [160].
We consider that whenever a mutant phenotype of D. melanogaster is presumably determined by a new mutant allele relevant for modeling an hGD, an intra-specific rescue with the wild-type copy of dGOI is recommended to confirm this link. There is always a potential risk that the mutant phenotype might have been induced by an unknown mutation, most probably residing on the same chromosome. In such case, the respective mutant strain is not appropriate for heterologous rescue experiments.
A positive heterologous rescue reveals that a basic interactome required for normal functions of the orthologous proteins was conserved between the two species. This aspect favors modeling of the disorder mainly when a chemical rescue is to be further attempted. Targeted mutagenesis of an appropriate dGOI often results in impaired fruit flies mimicking mutant phenotypes specific for the respective hGD. These mutants offer an excellent perspective for understanding the genetic mechanisms fueling the disorder since they are ideal to be subjected to chemical rescue assays, namely testing of candidate drugs on them. Nevertheless, care should be taken when the chemical rescue is efficient in mutant fruit flies impaired by disease-associated fly or human alleles without a preliminary successful heterologous rescue, since conclusions may be wrong. The candidate chemical may not target the mutant protein encoded by dGOILOF but rather a direct or a close interactor of it. Therefore, it was the adjustment of the stereo-chemical interaction with the defective protein that led to heterologous rescue instead of a repairing process. If the respective interactor is different or absent in mammals, the positive effects of the putative drug may not be reproducible in human patients.
Both a partial rescue and the rescue failure may be explained by either a less effective or an impaired activity of the associated human protein of interest (hPOI) in the context of fruit fly proteome [13]. When heterologous rescue fails, there are a few alternative explanations, and they should be considered before concluding that the two genes are indeed not functional orthologs. Most probably, the mammalian proximal interactome has changed during evolution, therefore that hGOI is per se unable to fit into a fruit fly genetic pathway. Therefore, the hGOI may be involved in similar biological processes in mammals but via different molecular avenues. This situation is expected to hinder the extrapolation of some promising chemical rescue results from fruit fly to a mammalian organism. Alternatively, hidden mutations in the genetic background of the fruit fly strain subjected to heterologous rescue may impede the functional complementation results. In this case, it is worthy to outcross the LOF allele in a new, isogenized background and repeat the rescue phenotype experiment. Finally, either the molecular construct used for embryo injection or the GAL4 driver may have inhibitory behaviors, hence alternative approaches may need to be tested [162].
However, there is also a reverse of the medal, namely when one is performing a promising heterologous rescue experiment and does not obtain the functional complementation of the mutant fruit fly. We further discuss various hypothetical scenarios for successful and failed heterologous experiments, just to emphasize the complexity of data interpretation when hGDs are modeled on D. melanogaster.

6.1. Issues When Modeling hGDs in D. melanogaster Regardless of Positive Heterologous Rescue Results

In order to consider a mammalian genetic disorder to be modeled in D. melanogaster, a preliminary bioinformatics analysis of both nucleotide and amino acid sequences should be performed. Some scenarios concerning stereo-chemical interactions among an hPOIWT and its proximal interacting proteins which are prone to affect the results of heterologous rescue experiments are summarized in Figure 2.
Let us presume that a structurally orthologous dGOI is identified in the D. melanogaster genome and the human and fruit fly proteins share a common functional domain D1 (Figure 2A). Particularly, there may be also present an unknown functional domain D2 in the human protein (Figure 2A), which is not yet characterized and therefore is not described in the specific databases. In addition, let us assume that research papers refer to D1 as being affected in many of the reported patients, therefore, a research project is launched to model the respective disease in D. melanogaster. Nevertheless, in some patients, both D1 and D2 may be distorted, but we expect that there is no medical focus on D2 yet.
To construct the model strains, the fruit flies are subjected to targeted mutagenesis of dGOI. Further genetic and molecular analysis inquiry may reveal that the mutant phenotype of adult fruit flies, which mirror medical conditions, are caused by a distorted D1, and the LOF flies are rescued by hGOIWT. Candidate chemicals are then tested on the impaired flies and one potential drug compensates the steric distortion of D1, resulting in a partial or complete phenotype rescue of the fruit flies. However, unexpectedly, when this putative drug is tested in mutant mammals, there is no phenotype rescue, because hPOIWT interacts with the equivalent human interacting protein 1 (hIP1) and hIP2 in a different manner in mammals, involving also distinct domains of D2 of hPOIWT and hIP1 (Figure 2C). The drug tested on D. melanogaster was able to chemically rescue only the functional domain D1, which is homologous with D1 of hPOIWT but not the conformation of D2. Therefore, a new project is required to understand functions of D2 and then to identify a different drug capable to rescue its spatial conformation.
A special situation is encountered when IP1, IP2 and IP3 are present in both flies and humans, but the hPOIWT/hIP1/hIP2 complex is organized by different interactions. The heterologous rescue should not work since, in humans, D1 of hPOIWT is required along with D3 and D4 domains to interact with hIP2 in order to ignite coupling of hIP3 (Figure 2D). Steric constraints caused by D4 prevent formation of hybrid hPOIWT/dIP1 and functional complementation fails, although the targeted D1 is conserved between H. sapiens and D. melanogaster (Figure 2D).
In this case, a preliminary, negative heterologous rescue test suggests that modeling of the respective genetic disorder is more appropriately performed in a mammalian experimental model such as mouse.

6.2. Possible Scenarios Accounting for Heterologous Rescue Failure or Partial Rescue in D. melanogaster

Let us consider that a protein complex dPOIWT (Drosophila POIWT)/dIP1 (Drosophila IP1) interacts with dIP2. The resulting complex dPOIWT/dIP1/dIP2 activates dIP3, which is a key prerequisite for the normal functioning of a biological process (Figure 2B). The orthologous gene employed for heterologous rescue of specific LOF fruit flies encodes for an hPOIWT, for which bioinformatics reveals the presence of the required domain that allows a proper interaction with dIP1. The particular stereochemistry of the hybrid hPOIWT/dIP1 complex, however, may result in a loose interconnection with dIP2 (Figure 2B). As a consequence, the hPOIWT/dIP1/dIP2 complex is unstable and even unable to activate dIP3. Therefore, the impaired biological process of LOF fruit flies is either partially rescued or not rescued at all by the human orthologous gene (Figure 2B).
The partial rescue results are still an incentive for modeling hGDs [13], but care should also be taken to the message send by an incomplete rescue result. Namely, does a partial rescue result resemble the case depicted in Figure 2B (bottom) or was the complete rescue phenotype impaired by genetic modifiers already present in the genetic background of a particular fruit fly strain? The roles of genetic modifiers as enhancers and suppressors in either precipitation or postponing the onset of symptoms should always be considered, as they represent the core of personalized medical genetics.
Complete heterologous rescue experiments are more encouraging when starting to model an hGD on D. melanogaster, but what about the ideal scenario when the orthologous gene of fruit fly rescues the mammalian mutant phenotypes? The orthologous gene pair atonal (ato, from fruit flies) and Math1 (from mouse) are both involved in the development of the nervous system of each species, and null alleles were reported for both genes. In an impressive experiment, transgenic Math1WT allele saved fruit flies with ato null phenotypes. Nonetheless, a transgenic atoWT allele copy also rescued lethality of Math1 null mice, otherwise unable to initiate breathing movement after birth; the heterologous rescued mice survived to adulthood [104].
Such examples of functional complementation, when mammalian or even other vertebrate orthologs save the fruit fly phenotypes, deserve a strong consideration if the respective gene is associated with an hGD.

7. Conclusions

Interdisciplinary research teaming up experts in genetics, bioinformatics, genomics and other medical domains strongly relies on D. melanogaster to model both mechanisms and treatment attempts of several impacting hGDs, such as neurological and cardiac disorders, cancers and infectious diseases.
Data emerging from functional complementation assays are very valuable for understanding the contribution of mutations associated with a genetic disorder in the context of an individual genetic background. In this context, carefully designed heterologous rescue experiments are a powerful tool to tackle defiant medical conditions via genetic avenues kept open by the D. melanogaster experimental model. The functional complementation assay is of prime relevance not only for medical use but also for fundamental research. What are these experiments telling us by revealing that many mammalian genes still function in the molecular context of the ancient D. melanogaster genome? Perhaps, this spectacular journey back in time evidences that numerous gene functions are conserved under an inherent selective pressure force: fundamental biological processes are articulated by using the same old molecular tools.
Obviously, some other medical challenges, such as aging issues, are prone to be modeled in D. melanogaster, but heterologous rescue assays’ data are still expected for specific orthologs genes. Although diseases such as Hutchinson–Gilford progeria syndrome are figured in fruit fly [284,285], we found no functional complementation experiments reported so far, neither in FlyBase nor in a recent review [286] for genes involved in aging pathology.
Hence, we conclude that preliminary heterologous rescue assays should serve as the standard for genetic analysis of hGDs on the D. melanogaster model.

Author Contributions

Conceptualization, A.A.E.; investigation, A.A.E., A.C.R., M.C.C. and M.M.M.; data curation, A.A.E. and A.C.R.; writing—original draft preparation, A.A.E., A.C.R., M.C.C. and M.M.M.; writing—review and editing, A.A.E. and A.C.R.; visualization, A.A.E., A.C.R. and M.C.C.; supervision, A.A.E.; funding acquisition, M.C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the C1.2.PFE-CDI.2021-587 project. This project is financed by the Ministry of Research, Innovation and Digitalization through Program 1—Development of the national R&D system, Subprogram 1.2—Institutional performance—Financing projects for excellence in RDI, Contract no. 41 PFE/30.12.2021. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript or in the decision to publish the results.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. De Maria Marchiano, R.; Di Sante, G.; Piro, G.; Carbone, C.; Tortora, G.; Boldrini, L.; Pietragalla, A.; Daniele, G.; Tredicine, M.; Cesario, A.; et al. Translational research in the era of precision medicine: Where we are and where we will go. J. Pers. Med. 2021, 11, 216. [Google Scholar] [CrossRef] [PubMed]
  2. Beckingham, K.M.; Armstrong, J.D.; Texada, M.J.; Munjaal, R.; Baker, D.A. Drosophila melanogaster—The model organism of choice for the complex biology of multi-cellular organisms. Gravit. Space Biol. Bull. Publ. Am. Soc. Gravit. Space Biol. 2005, 18, 17–29. [Google Scholar]
  3. Adams, M.D.; Celniker, S.E.; Holt, R.A.; Evans, C.A.; Gocayne, J.D.; Amanatides, P.G.; Scherer, S.E.; Li, P.W.; Hoskins, R.A.; Galle, R.F.; et al. The genome sequence of Drosophila melanogaster. Science 2000, 287, 2185–2195. [Google Scholar] [CrossRef] [Green Version]
  4. Venter, J.C.; Adams, M.D.; Myers, E.W.; Li, P.W.; Mural, R.J.; Sutton, G.G.; Smith, H.O.; Yandell, M.; Evans, C.A.; Holt, R.A.; et al. The sequence of the human genome. Science 2001, 291, 1304–1351. [Google Scholar] [CrossRef] [Green Version]
  5. Lander, E.S.; Linton, L.M.; Birren, B.; Nusbaum, C.; Zody, M.C.; Baldwin, J.; Devon, K.; Dewar, K.; Doyle, M.; FitzHugh, W.; et al. Initial sequencing and analysis of the human genome. Nature 2001, 409, 860–921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Dunham, M.J.; Fowler, D.M. Contemporary, yeast-based approaches to understanding human genetic variation. Curr. Opin. Genet. Dev. 2013, 23, 658–664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Cherry, J.M.; Hong, E.L.; Amundsen, C.; Balakrishnan, R.; Binkley, G.; Chan, E.T.; Christie, K.R.; Costanzo, M.C.; Dwight, S.S.; Engel, S.R.; et al. Saccharomyces Genome Database: The genomics resource of budding yeast. Nucleic Acids Res. 2012, 40, D700–D705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Larkin, A.; Marygold, S.J.; Antonazzo, G.; Attrill, H.; Dos Santos, G.; Garapati, P.V.; Goodman, J.L.; Gramates, L.S.; Millburn, G.; Strelets, V.B.; et al. FlyBase: Updates to the Drosophila melanogaster knowledge base. Nucleic Acids Res. 2021, 49, D899–D907. [Google Scholar] [CrossRef]
  9. Bellen, H.J.; Wangler, M.F.; Yamamoto, S. The fruit fly at the interface of diagnosis and pathogenic mechanisms of rare and common human diseases. Hum. Mol. Genet. 2019, 28, R207–R214. [Google Scholar] [CrossRef] [Green Version]
  10. Specchia, V.; Puricella, A.; D’Attis, S.; Massari, S.; Giangrande, A.; Bozzetti, M.P. Drosophila melanogaster as a Model to Study the Multiple Phenotypes, Related to Genome Stability of the Fragile-X Syndrome. Front. Genet. 2019, 10, 10. [Google Scholar] [CrossRef]
  11. Bellen, H.J.; Yamamoto, S. Morgan’s legacy: Fruit flies and the functional annotation of conserved genes. Cell 2015, 163, 12–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Rorth, P. Gal4 in the Drosophila female germline. Mech. Dev. 1998, 78, 113–118. [Google Scholar] [CrossRef] [Green Version]
  13. Wangler, M.F.; Yamamoto, S.; Chao, H.-T.; Posey, J.E.; Westerfield, M.; Postlethwait, J.; Members of the Undiagnosed Diseases Network (UDN); Hieter, P.; Boycott, K.M.; Campeau, P.M.; et al. Model Organisms Facilitate Rare Disease Diagnosis and Therapeutic Research. Genetics 2017, 207, 9–27. [Google Scholar] [CrossRef]
  14. Harnish, J.M.; Deal, S.L.; Chao, H.T.; Wangler, M.F.; Yamamoto, S. In Vivo Functional Study of Disease-associated Rare Human Variants Using Drosophila. J. Vis. Exp. JoVE 2019, 150, e59658. [Google Scholar] [CrossRef] [Green Version]
  15. Lu, C.; Vihtelic, T.S.; Hyde, D.R.; Li, T. A neuronal-specific mammalian homolog of the Drosophila retinal degeneration B gene with expression restricted to the retina and dentate gyrus. J. Neurosci. Off. J. Soc. Neurosci. 1999, 19, 7317–7325. [Google Scholar] [CrossRef] [Green Version]
  16. Leiserson, W.M.; Harkins, E.W.; Keshishian, H. Fray, a Drosophila serine/threonine kinase homologous to mammalian PASK, is required for axonal ensheathment. Neuron 2000, 28, 793–806. [Google Scholar] [CrossRef] [Green Version]
  17. Maurya, B.; Surabhi, S.; Pandey, P.; Mukherjee, A.; Mutsuddi, M. Insights into Human Neurodegeneration: Lessons Learnt from Drosophila; Springer: New York, NY, USA, 2019; pp. 373–403. [Google Scholar] [CrossRef]
  18. Qiao, H.H.; Wang, F.; Xu, R.G.; Sun, J.; Zhu, R.; Mao, D.; Ren, X.; Wang, X.; Jia, Y.; Peng, P.; et al. An efficient and multiple target transgenic RNAi technique with low toxicity in Drosophila. Nat. Commun. 2018, 9, 4160. [Google Scholar] [CrossRef] [Green Version]
  19. McGurk, L.; Berson, A.; Bonini, N.M. Drosophila as an In Vivo Model for Human Neurodegenerative Disease. Genetics 2015, 201, 377–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Lloyd, T.E.; Taylor, J.P. Flightless flies: Drosophila models of neuromuscular disease. Ann. N. Y. Acad. Sci. 2010, 1184, e1–e20. [Google Scholar] [CrossRef] [Green Version]
  21. Rubin, G.M.; Yandell, M.D.; Wortman, J.R.; Gabor Miklos, G.L.; Nelson, C.R.; Hariharan, I.K.; Fortini, M.E.; Li, P.W.; Apweiler, R.; Fleischmann, W.; et al. Comparative genomics of the eukaryotes. Science 2000, 287, 2204–2215. [Google Scholar] [CrossRef] [Green Version]
  22. Brody, T. The Interactive Fly: Gene networks, development and the Internet. Trends Genet. 1999, 15, 333–334. [Google Scholar] [CrossRef]
  23. Yang, Y.; Nishimura, I.; Imai, Y.; Takahashi, R.; Lu, B. Parkin suppresses dopaminergic neuronselective neurotoxicity induced by Pael-R in Drosophila. Neuron 2003, 37, 911–924. [Google Scholar] [CrossRef] [Green Version]
  24. Burchell, V.S.; Nelson, D.E.; Sanchez-Martinez, A.; Delgado-Camprubi, M.; Ivatt, R.M.; Pogson, J.H.; Randle, S.J.; Wray, S.; Lewis, P.A.; Houlden, H.; et al. The Parkinson’s disease–linked proteins Fbxo7 and Parkin interact to mediate mitophagy. Nat. Neurosci. 2013, 16, 1257–1265. [Google Scholar] [CrossRef] [Green Version]
  25. Paisán-Ruiz, C.; Guevara, R.; Federoff, M.; Hanagasi, H.; Sina, F.; Elahi, E.; Schneider, S.A.; Schwingenschuh, P.; Bajaj, N.; Emre, M.; et al. Early-onset L-dopa-responsive parkinsonism with pyramidal signs due to ATP13A2, PLA2G6, FBXO7 and spatacsin mutations. Mov. Disord. 2010, 25, 1791–1800. [Google Scholar] [CrossRef] [PubMed]
  26. Baughman, J.M.; Nilsson, R.; Gohil, V.M.; Arlow, D.H.; Gauhar, Z.; Mootha, V.K. A computational screen for regulators of oxidative phosphorylation implicates SLIRP in mitochondrial RNA homeostasis. PLoS Genet. 2009, 5, e1000590. [Google Scholar] [CrossRef] [Green Version]
  27. Meng, H.; Yamashita, C.; Shiba-Fukushima, K.; Inoshita, T.; Funayama, M.; Sato, S.; Hatta, T.; Natsume, T.; Umitsu, M.; Takagi, J.; et al. Loss of Parkinson’s disease-associated protein CHCHD2 affects mitochondrial crista structure and destabilizes cytochrome c. Nat. Commun. 2017, 8, 15500. [Google Scholar] [CrossRef] [Green Version]
  28. Mori, A.; Hatano, T.; Inoshita, T.; Shiba-Fukushima, K.; Koinuma, T.; Meng, H.; Kubo, S.-I.; Spratt, S.; Cui, C.; Yamashita, C.; et al. Parkinson’s disease-associated iPLA2-VIA/PLA2G6 regulates neuronal functions and α-synuclein stability through membrane remodeling. Proc. Natl. Acad. Sci. USA 2019, 116, 20689–20699. [Google Scholar] [CrossRef] [Green Version]
  29. Zhang, B.; Egli, D.; Georgiev, O.; Schaffner, W. The Drosophila homolog of mammalian zinc finger factor MTF-1 activates transcription in response to heavy metals. Mol. Cell. Biol. 2001, 21, 4505–4514. [Google Scholar] [CrossRef] [Green Version]
  30. Ghoshal, K.; Majumder, S.; Zhu, Q.; Hunzeker, J.; Datta, J.; Shah, M.; Sheridan, J.F.; Jacob, S.T. Influenza virus infection induces metallothionein gene expression in the mouse liver and lung by overlapping but distinct molecular mechanisms. Mol. Cell. Biol. 2001, 21, 8301–8317. [Google Scholar] [CrossRef] [Green Version]
  31. Saini, N.; Georgiev, O.; Schaffner, W. The parkin mutant phenotype in the fly is largely rescued by metal-responsive transcription factor (MTF-1). Mol. Cell. Biol. 2011, 31, 2151–2161. [Google Scholar] [CrossRef] [Green Version]
  32. Balamurugan, K.; Egli, D.; Selvaraj, A.; Zhang, B.; Georgiev, O.; Schaffner, W. Metal-responsive transcription factor (MTF-1) and heavy metal stress response in Drosophila and mammalian cells: A functional comparison. Biol. Chem. 2001, 385, 597–603. [Google Scholar] [CrossRef] [PubMed]
  33. Morais, V.A.; Verstreken, P.; Roethig, A.; Smet, J.; Snellinx, A.; Vanbrabant, M.; Haddad, D.; Frezza, C.; Mandemakers, W.; Vogt-Weisenhorn, D.; et al. Parkinson’s disease mutations in PINK1 result in decreased Complex I activity and deficient synaptic function. EMBO Mol. Med. 2009, 1, 99–111. [Google Scholar] [CrossRef]
  34. Brown, R.H.; Al-Chalabi, A. Amyotrophic Lateral Sclerosis. N. Engl. J. Med. 2017, 377, 162–172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Longinetti, E.; Fang, F. Epidemiology of amyotrophic lateral sclerosis: An update of recent literature. Curr. Opin. Neurol. 2019, 32, 771–776. [Google Scholar] [CrossRef] [PubMed]
  36. Abramzon, Y.A.; Fratta, P.; Traynor, B.J.; Chia, R. The Overlapping Genetics of Amyotrophic Lateral Sclerosis and Frontotemporal Dementia. Front. Neurosci. 2020, 14, 42. [Google Scholar] [CrossRef] [Green Version]
  37. Elamin, M.; Bede, P.; Byrne, S.; Jordan, N.; Gallagher, L.; Wynne, B.; O’Brien, C.; Phukan, J.; Lynch, C.; Pender, N.; et al. Cognitive changes predict functional decline in ALS: A population-based longitudinal study. Neurology 2013, 80, 1590–1597. [Google Scholar] [CrossRef]
  38. Layalle, S.; They, L.; Ourghani, S.; Raoul, C.; Soustelle, L. Amyotrophic Lateral Sclerosis Genes in Drosophila melanogaster. Int. J. Mol. Sci. 2021, 22, 904. [Google Scholar] [CrossRef]
  39. Picher-Martel, V.; Valdmanis, P.N.; Gould, P.V.; Julien, J.P.; Dupre, N. From animal models to human disease: A genetic approach for personalized medicine in ALS. Acta Neuropathol. Commun. 2016, 4, 70. [Google Scholar] [CrossRef]
  40. Liguori, F.; Amadio, S.; Volonté, C. Fly for ALS: Drosophila modeling on the route to amyotrophic lateral sclerosis modifiers. Cell. Mol. Life Sci. 2021, 78, 6143–6160. [Google Scholar] [CrossRef]
  41. Mockett, R.J.; Radyuk, S.N.; Benes, J.J.; Orr, W.C.; Sohal, R.S. Phenotypic effects of familial amyotrophic lateral sclerosis mutant Sod alleles in transgenic Drosophila. Proc. Natl. Acad. Sci. USA 2003, 100, 301–306. [Google Scholar] [CrossRef] [Green Version]
  42. Parkes, T.L.; Elia, A.J.; Dickinson, D.; Hilliker, A.J.; Phillips, J.P.; Boulianne, G.L. Extension of Drosophila lifespan by overexpression of human SOD1 in motorneurons. Nat. Genet. 1998, 19, 171–174. [Google Scholar] [CrossRef] [PubMed]
  43. Tollervey, J.R.; Curk, T.; Rogelj, B.; Briese, M.; Cereda, M.; Kayikci, M.; Konig, J.; Hortobagyi, T.; Nishimura, A.L.; Zupunski, V.; et al. Characterizing the RNA targets and position-dependent splicing regulation by TDP-43. Nat. Neurosci. 2011, 14, 452–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Alami, N.H.; Smith, R.B.; Carrasco, M.A.; Williams, L.A.; Winborn, C.S.; Han, S.S.; Kiskinis, E.; Winborn, B.; Freibaum, B.D.; Kanagaraj, A.; et al. Axonal transport of TDP-43 mRNA granules is impaired by ALS-causing mutations. Neuron 2014, 81, 536–543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Honda, D.; Ishigaki, S.; Iguchi, Y.; Fujioka, Y.; Udagawa, T.; Masuda, A.; Ohno, K.; Katsuno, M.; Sobue, G. The ALS/FTLD-related RNA-binding proteins TDP-43 and FUS have common downstream RNA targets in cortical neurons. FEBS Open Bio 2013, 4, 1–10. [Google Scholar] [CrossRef] [Green Version]
  46. Baldwin, K.R.; Godena, V.K.; Hewitt, V.L.; Whitworth, A.J. Axonal transport defects are a common phenotype in Drosophila models of ALS. Hum. Mol. Genet. 2016, 25, 2378–2392. [Google Scholar] [CrossRef] [Green Version]
  47. Wang, J.W.; Brent, J.R.; Tomlinson, A.; Shneider, N.A.; McCabe, B.D. The ALS-associated proteins FUS and TDP-43 function together to affect Drosophila locomotion and life span. J. Clin. Investig. 2011, 121, 4118–4126. [Google Scholar] [CrossRef] [Green Version]
  48. Arlington, V.A. Diagnostic and Statistical Manual of Mental Disorders, 5th ed.; The American Psychiatric Association: Washington, DC, USA, 2013. [Google Scholar]
  49. Lord, C.; Brugha, T.S.; Charman, T.; Cusack, J.; Dumas, G.; Frazier, T.; Jones, E.J.H.; Jones, R.M.; Pickles, A.; State, M.W.; et al. Autism spectrum disorder. Nat. Rev. Dis. Prim. 2020, 6, 5. [Google Scholar] [CrossRef]
  50. Iovene, M.R.; Bombace, F.; Maresca, R.; Sapone, A.; Iardino, P.; Picardi, A.; Marotta, R.; Schiraldi, C.; Siniscalco, D.; Serra, N.; et al. Intestinal Dysbiosis and Yeast Isolation in Stool of Subjects with Autism Spectrum Disorders. Mycopathologia 2016, 182, 349–363. [Google Scholar] [CrossRef] [Green Version]
  51. Fulceri, F.; Morelli, M.; Santocchi, E.; Cena, H.; Del Bianco, T.; Narzisi, A.; Calderoni, S.; Muratori, F. Gastrointestinal symptoms and behavioral problems in preschoolers with autism spectrum disorder. Dig. Liver Dis. 2016, 48, 248–254. [Google Scholar] [CrossRef]
  52. Marler, S.; Ferguson, B.J.; Lee, E.B.; Peters, B.; Williams, K.C.; McDonnell, E.; Macklin, E.A.; Levitt, P.; Margolis, K.G.; Beversdorf, D.Q.; et al. Association of rigid-compulsive behavior with functional constipation in autism spectrum disorder. J. Autism Dev. Disord. 2017, 47, 1673–1681. [Google Scholar] [CrossRef]
  53. Abrahams, B.S.; Arking, D.E.; Campbell, D.B.; Mefford, H.C.; Morrow, E.M.; Weiss, L.A.; Menashe, I.; Wadkins, T.; Banerjee-Basu, S.; Packer, A. SFARI Gene 2.0: A community-driven knowledgebase for the autism spectrum disorders (ASDs). Mol. Autism 2013, 4, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Satterstrom, F.K.; Kosmicki, J.A.; Wang, J.; Breen, M.S.; De Rubeis, S.; An, J.Y.; Peng, M.; Collins, R.; Grove, J.; Klei, L.; et al. Large-Scale Exome Sequencing Study Implicates Both Developmental and Functional Changes in the Neurobiology of Autism. Cell 2020, 180, 568–584.e23. [Google Scholar] [CrossRef] [PubMed]
  55. Salim, S.; Banu, A.; Alwa, A.; Gowda, S.B.M.; Mohammad, F. The gut-microbiota-brain axis in autism: What Drosophila models can offer? J. Neurodev. Disord. 2021, 13, 37. [Google Scholar] [CrossRef]
  56. Hu, Y.; Flockhart, I.; Vinayagam, A.; Bergwitz, C.; Berger, B.; Perrimon, N.; Mohr, S.E. An integrative approach to ortholog prediction for disease-focused and other functional studies. BMC Bioinform. 2011, 12, 357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Clifford, S.; Dissanayake, C.; Bui, Q.M.; Huggins, R.; Taylor, A.K.; Loesch, D.Z. Autism spectrum phenotype in males and females with fragile X full mutation and premutation. J. Autism Dev. Disord. 2006, 37, 738–747. [Google Scholar] [CrossRef]
  58. Coffee, R.L., Jr.; Tessier, C.R.; Woodruff, E.A., 3rd; Broadie, K. Fragile X mental retardation protein has a unique, evolutionarily conserved neuronal function not shared with FXR1P or FXR2P. Dis. Model. Mech. 2010, 3, 471–485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Li, J.; Yen, C.; Liaw, D.; Podsypanina, K.; Bose, S.; Wang, S.I.; Puc, J.; Miliaresis, C.; Rodgers, L.; McCombie, R.; et al. PTEN, a putative protein tyrosine phosphatase gene mutated in human brain, breast, and prostate cancer. Science 1997, 275, 1943–1948. [Google Scholar] [CrossRef] [PubMed]
  60. Butler, M.G.; Dasouki, M.J.; Zhou, X.P.; Talebizadeh, Z.; Brown, M.; Takahashi, T.N.; Miles, J.H.; Wang, C.H.; Stratton, R.; Pilarski, R.; et al. Subset of individuals with autism spectrum disorders and extreme macrocephaly associated with germline PTEN tumour suppressor gene mutations. J. Med. Genet. 2005, 42, 318–321. [Google Scholar] [CrossRef] [Green Version]
  61. Mester, J.L.; Ghosh, R.; Pesaran, T.; Huether, R.; Karam, R.; Hruska, K.S.; Costa, H.A.; Lachlan, K.; Ngeow, J.; Barnholtz-Sloan, J.; et al. Gene-specific criteria for PTEN variant curation: Recommendations from the ClinGen PTEN Expert Panel. Hum. Mutat. 2018, 39, 1581–1592. [Google Scholar] [CrossRef]
  62. Sun, H.; Lesche, R.; Li, D.M.; Liliental, J.; Zhang, H.; Gao, J.; Gavrilova, N.; Mueller, B.; Liu, X.; Wu, H. PTEN modulates cell cycle progression and cell survival by regulating phosphatidylinositol 3,4,5,-trisphosphate and Akt/protein kinase B signaling pathway. Proc. Natl. Acad. Sci. USA 1999, 96, 6199–6204. [Google Scholar] [CrossRef] [Green Version]
  63. Ganguly, P.; Madonsela, L.; Chao, J.T.; Loewen, C.J.R.; O’Connor, T.P.; Verheyen, E.M.; Allan, D.W. A scalable Drosophila assay for clinical interpretation of human PTEN variants in suppression of PI3K/AKT induced cellular proliferation. PLoS Genet. 2021, 17, e1009774. [Google Scholar] [CrossRef]
  64. Pelc, K.; Cheron, G.; Dan, B. Behavior and neuropsychiatric manifestations in Angelman syndrome. Neuropsychiatr. Dis. Treat. 2008, 4, 577–584. [Google Scholar] [PubMed] [Green Version]
  65. Ratiu, A.C.; Ecovoiu, A.A.; Graur, M.; Gavrila, L. A second site lethal mutation masked the real phenotype of EP(3)3214 transgenic line. Bull. USAMV Anim. Sci. Biotechnol. 2008, 65, 475. [Google Scholar]
  66. Ratiu, A.C.; Neagu, A.; Mihalache, M.R.; Lazar, V. Long-term administration of omega-3 fatty acids alleviates Angelman syndrome-like phenotype in an Ube3a mutant strain of Drosophila melanogaster. Biointerface Res. Appl. Chem. 2015, 5, 996–1002. [Google Scholar]
  67. Chakraborty, M.; Paul, B.K.; Nayak, T.; Das, A.; Jana, N.R.; Bhutani, S. The E3 ligase ube3a is required for learning in Drosophila melanogaster. Biochem. Biophys. Res. Commun. 2015, 462, 71–77. [Google Scholar] [CrossRef]
  68. Chai, A.; Withers, J.; Koh, Y.H.; Parry, K.; Bao, H.; Zhang, B.; Budnik, V.; Pennetta, G. hVAPB, the causative gene of a heterogeneous group of motor neuron diseases in humans, is functionally interchangeable with its Drosophila homologue DVAP-33A at the neuromuscular junction. Hum. Mol. Genet. 2007, 17, 266–280. [Google Scholar] [CrossRef] [Green Version]
  69. Besson, M.T.; Dupont, P.; Fridell, Y.W.; Liévens, J.C. Increased energy metabolism rescues glia-induced pathology in a Drosophila model of Huntington’s disease. Hum. Mol. Genet. 2010, 19, 3372–3382. [Google Scholar] [CrossRef] [Green Version]
  70. Rui, Y.N.; Xu, Z.; Patel, B.; Chen, Z.; Chen, D.; Tito, A.; David, G.; Sun, Y.; Stimming, E.F.; Bellen, H.J.; et al. Huntingtin functions as a scaffold for selective macroautophagy. Nat. Cell Biol. 2015, 17, 262–275. [Google Scholar] [CrossRef] [Green Version]
  71. Tsai, P.I.; Lin, C.H.; Hsieh, C.H.; Papakyrikos, A.M.; Kim, M.J.; Napolioni, V.; Schoor, C.; Couthouis, J.; Wu, R.M.; Wszolek, Z.K.; et al. PINK1 Phosphorylates MIC60/Mitofilin to Control Structural Plasticity of Mitochondrial Crista Junctions. Mol. Cell 2018, 69, 744–756.e6. [Google Scholar] [CrossRef] [Green Version]
  72. Poole, A.C.; Thomas, R.E.; Andrews, L.A.; McBride, H.M.; Whitworth, A.J.; Pallanck, L.J. The PINK1/Parkin pathway regulates mitochondrial morphology. Proc. Natl. Acad. Sci. USA 2008, 105, 1638–1643. [Google Scholar] [CrossRef] [Green Version]
  73. Wang, C.; Lu, R.; Ouyang, X.; Ho, M.W.; Chia, W.; Yu, F.; Lim, K.L. Drosophila overexpressing parkin R275W mutant exhibits dopaminergic neuron degeneration and mitochondrial abnormalities. J. Neurosci. Off. J. Soc. Neurosci. 2007, 27, 8563–8570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Soukup, S.F.; Kuenen, S.; Vanhauwaert, R.; Manetsberger, J.; Hernández-Díaz, S.; Swerts, J.; Schoovaerts, N.; Vilain, S.; Gounko, N.V.; Vints, K.; et al. LRRK2-Dependent EndophilinA Phosphoswitch Is Critical for Macroautophagy at Presynaptic Terminals. Neuron 2016, 92, 829–844. [Google Scholar] [CrossRef] [Green Version]
  75. Chuang, C.-L.; Lu, Y.-N.; Wang, H.-C.; Chang, H.-Y. Genetic dissection reveals that Akt is the critical kinase downstream of LRRK2 to phosphorylate and inhibit FOXO1, and promotes neuron survival. Hum. Mol. Genet. 2014, 23, 5649–5658. [Google Scholar] [CrossRef] [PubMed]
  76. Li, B.; Wong, C.; Gao, S.M.; Zhang, R.; Sun, R.; Li, Y.; Song, Y. The retromer complex safeguards against neural progenitor-derived tumorigenesis by regulating Notch receptor trafficking. eLife 2018, 7, e38181. [Google Scholar] [CrossRef] [PubMed]
  77. Bolkan, B.J.; Kretzschmar, D. Loss of Tau results in defects in photoreceptor development and progressive neuronal degeneration in Drosophila. Dev. Neurobiol. 2014, 74, 1210–1225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Wu, C.-H.; Giampetruzzi, A.; Tran, H.; Fallini, C.; Gao, F.-B.; Landers, J.E. A Drosophila model of ALS reveals a partial loss of function of causative human PFN1 mutants. Hum. Mol. Genet. 2017, 26, 2146–2155. [Google Scholar] [CrossRef] [Green Version]
  79. Johnson, A.E.; Shu, H.; Hauswirth, A.G.; Tong, A.; Davis, G.W. VCP-dependent muscle degeneration is linked to defects in a dynamic tubular lysosomal network in vivo. eLife 2015, 4, e07366. [Google Scholar] [CrossRef]
  80. Jakobsdottir, J.; van der Lee, S.J.; Bis, J.C.; Chouraki, V.; Li-Kroeger, D.; Yamamoto, S.; Grove, M.L.; Naj, A.; Vronskaya, M.; Salazar, J.L.; et al. Rare Functional Variant in TM2D3 is Associated with Late-Onset Alzheimer’s Disease. PLoS Genet. 2016, 12, e1006327, Erratum in PLoS Genet. 2016, 12, e1006456. [Google Scholar] [CrossRef] [Green Version]
  81. Buechling, T.; Bartscherer, K.; Ohkawara, B.; Chaudhary, V.; Spirohn, K.; Niehrs, C.; Boutros, M. Wnt/Frizzled signaling requires dPRR, the Drosophila homolog of the prorenin receptor. Curr. Biol. 2010, 20, 1263–1268. [Google Scholar] [CrossRef] [Green Version]
  82. Ghosh, S.G.; Becker, K.; Huang, H.; Dixon-Salazar, T.; Chai, G.; Salpietro, V.; Al-Gazali, L.; Waisfisz, Q.; Wang, H.; Vaux, K.K.; et al. Biallelic Mutations in ADPRHL2, Encoding ADP-Ribosylhydrolase 3, Lead to a Degenerative Pediatric Stress-Induced Epileptic Ataxia Syndrome. Am. J. Hum. Genet. 2018, 103, 431–439. [Google Scholar] [CrossRef] [Green Version]
  83. Broeck, L.V.; Kleinberger, G.; Chapuis, J.; Gistelinck, M.; Amouyel, P.; Van Broeckhoven, C.; Lambert, J.-C.; Callaerts, P.; Dermaut, B. Functional complementation in Drosophila to predict the pathogenicity of TARDBP variants: Evidence for a loss-of-function mechanism. Neurobiol. Aging 2014, 36, 1121–1129. [Google Scholar] [CrossRef] [PubMed]
  84. Dutta, D.; Briere, L.C.; Kanca, O.; Marcogliese, P.; Walker, M.A.; High, F.A.; Vanderver, A.; Krier, J.; Carmichael, N.; Callahan, C.; et al. De novo mutations in TOMM70, a receptor of the mitochondrial import translocase, cause neurological impairment. Hum. Mol. Genet. 2020, 29, 1568–1579. [Google Scholar] [CrossRef] [PubMed]
  85. Sunderhaus, E.R.; Law, A.D.; Kretzschmar, D. Disease-Associated PNPLA6 Mutations Maintain Partial Functions When Analyzed in Drosophila. Front. Neurosci. 2019, 13, 1207. [Google Scholar] [CrossRef] [PubMed]
  86. Topaloglu, A.K.; Lomniczi, A.; Kretzschmar, D.; Dissen, G.A.; Kotan, L.D.; McArdle, C.A.; Koc, A.F.; Hamel, B.C.; Guclu, M.; Papatya, E.D.; et al. Loss-of-Function Mutations in PNPLA6 Encoding Neuropathy Target Esterase Underlie Pubertal Failure and Neurological Deficits in Gordon Holmes Syndrome. J. Clin. Endocrinol. Metab. 2014, 99, E2067–E2075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Yang, Y.; Wu, Z.; Kuo, Y.M.; Zhou, B. Dietary rescue of fumble–a Drosophila model for pantothenate-kinase-associated neurodegeneration. J. Inherit. Metab. Dis. 2005, 28, 1055–1064. [Google Scholar] [CrossRef]
  88. Ohno, M.; Hiraoka, Y.; Matsuoka, T.; Tomimoto, H.; Takao, K.; Miyakawa, T.; Oshima, N.; Kiyonari, H.; Kimura, T.; Kita, T.; et al. Nardilysin regulates axonal maturation and myelination in the central and peripheral nervous system. Nat. Neurosci. 2009, 12, 1506–1513. [Google Scholar] [CrossRef] [Green Version]
  89. Yoon, W.H.; Sandoval, H.; Nagarkar-Jaiswal, S.; Jaiswal, M.; Yamamoto, S.; Haelterman, N.A.; Putluri, N.; Putluri, V.; Sreekumar, A.; Tos, T.; et al. Loss of Nardilysin, a Mitochondrial Co-chaperone for α-Ketoglutarate Dehydrogenase, Promotes mTORC1 Activation and Neurodegeneration. Neuron 2016, 93, 115–131. [Google Scholar] [CrossRef] [Green Version]
  90. Vonk, J.J.; Yeshaw, W.M.; Pinto, F.; Faber, A.I.; Lahaye, L.L.; Kanon, B.; van der Zwaag, M.; Velayos-Baeza, A.; Freire, R.; van IJzendoorn, S.C.; et al. Drosophila Vps13 Is Required for Protein Homeostasis in the Brain. PLoS ONE 2017, 12, e0170106. [Google Scholar] [CrossRef]
  91. Yeshaw, W.M.; van der Zwaag, M.; Pinto, F.; Lahaye, L.L.; Faber, A.I.; Gómez-Sánchez, R.; Dolga, A.M.; Poland, C.; Monaco, A.P.; van IJzendoorn, S.C.; et al. Human VPS13A is associated with multiple organelles and influences mitochondrial morphology and lipid droplet motility. eLife 2019, 8, e43561. [Google Scholar] [CrossRef]
  92. Xiong, B.; Bayat, V.; Jaiswal, M.; Zhang, K.; Sandoval, H.; Charng, W.L.; Li, T.; David, G.; Duraine, L.; Lin, Y.Q.; et al. Crag Is a GEF for Rab11 required for rhodopsin trafficking and maintenance of adult photoreceptor cells. PLOS Biol. 2012, 10, e1001438. [Google Scholar] [CrossRef] [Green Version]
  93. Miyake, N.; Fukai, R.; Ohba, C.; Chihara, T.; Miura, M.; Shimizu, H.; Kakita, A.; Imagawa, E.; Shiina, M.; Ogata, K.; et al. Biallelic TBCD Mutations Cause Early-Onset Neurodegenerative Encephalopathy. Am. J. Hum. Genet. 2016, 99, 950–961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Hekmat-Scafe, D.S.; Mercado, A.; Fajilan, A.A.; Lee, A.W.; Hsu, R.; Mount, D.B.; Tanouye, M.A. Seizure Sensitivity Is Ameliorated by Targeted Expression of K+–Cl− Cotransporter Function in the Mushroom Body of the Drosophila Brain. Genetics 2010, 184, 171–183. [Google Scholar] [CrossRef] [Green Version]
  95. Praschberger, R.; Lowe, S.A.; Malintan, N.T.; Giachello, C.; Patel, N.; Houlden, H.; Kullmann, D.M.; Baines, R.A.; Usowicz, M.M.; Krishnakumar, S.S.; et al. Mutations in Membrin/GOSR2 Reveal Stringent Secretory Pathway Demands of Dendritic Growth and Synaptic Integrity. Cell Rep. 2017, 21, 97–109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Tea, J.S.; Luo, L. The chromatin remodeling factor Bap55 functions through the TIP60 complex to regulate olfactory projection neuron dendrite targeting. Neural Dev. 2011, 6, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kunduri, G.; Turner-Evans, D.; Konya, Y.; Izumi, Y.; Nagashima, K.; Lockett, S.; Holthuis, J.; Bamba, T.; Acharya, U.; Acharya, J.K. Defective cortex glia plasma membrane structure underlies light-induced epilepsy in cpes mutants. Proc. Natl. Acad. Sci. USA 2018, 115, E8919. [Google Scholar] [CrossRef] [Green Version]
  98. Tenedini, F.M.; Saez Gonzalez, M.; Hu, C.; Pedersen, L.H.; Petruzzi, M.M.; Spitzweck, B.; Wang, D.; Richter, M.; Petersen, M.; Szpotowicz, E.; et al. Maintenance of cell type-specific connectivity and circuit function requires Tao kinase. Nat. Commun. 2019, 10, 3506. [Google Scholar] [CrossRef] [Green Version]
  99. Hu, C.; Kanellopoulos, A.; Richter, M.; Petersen, M.; Konietzny, A.; Tenedini, F.M.; Hoyer, N.; Cheng, L.; Poon, C.L.; Harvey, K.F.; et al. Conserved Tao Kinase Activity Regulates Dendritic Arborization, Cytoskeletal Dynamics, and Sensory Function in Drosophila. J. Neurosci. 2020, 40, 1819–1833. [Google Scholar] [CrossRef]
  100. Hamilton, P.J.; Campbell, N.G.; Sharma, S.; Erreger, K.; Herborg Hansen, F.; Saunders, C.; Belovich, A.N.; NIH ARRA Autism Sequencing Consortium; Sahai, M.A.; Cook, E.H.; et al. De novo mutation in the dopamine transporter gene associates dopamine dysfunction with autism spectrum disorder. Mol. Psychiatry 2013, 18, 1315–1323. [Google Scholar] [CrossRef] [Green Version]
  101. Campbell, N.G.; Shekar, A.; Aguilar, J.I.; Peng, D.; Navratna, V.; Yang, D.; Morley, A.N.; Duran, A.M.; Galli, G.; O’Grady, B.; et al. Structural, functional, and behavioral insights of dopamine dysfunction revealed by a deletion in SLC6A3. Proc. Natl. Acad. Sci. USA 2019, 116, 3853–3862. [Google Scholar] [CrossRef] [Green Version]
  102. Zheng, J.C.; Tham, C.T.; Keatings, K.; Fan, S.; Liou, A.Y.-C.; Numata, Y.; Allan, D.; Numata, M. Secretory Carrier Membrane Protein (SCAMP) deficiency influences behavior of adult flies. Front. Cell Dev. Biol. 2014, 2, 64. [Google Scholar] [CrossRef] [Green Version]
  103. Volders, K.; Scholz, S.; Slabbaert, J.R.; Nagel, A.C.; Verstreken, P.; Creemers, J.W.; Callaerts, P.; Schwärzel, M. Drosophila rugose is a functional homolog of mammalian Neurobeachin and affects synaptic architecture, brain morphology, and associative learning. J. Neurosci. 2012, 32, 15193–15204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Wang, V.Y.; Hassan, B.A.; Bellen, H.J.; Zoghbi, H.Y. Drosophila atonal fully rescues the phenotype of Math1 null mice: New functions evolve in new cellular contexts. Curr Biol 2002, 12, 1611–1616. [Google Scholar] [CrossRef] [Green Version]
  105. Züchner, S.; Mersiyanova, I.V.; Muglia, M.; Bissar-Tadmouri, N.; Rochelle, J.; Dadali, E.L.; Zappia, M.; Nelis, E.; Patitucci, A.; Senderek, J.; et al. Mutations in the mitochondrial GTPase mitofusin 2 cause Charcot-Marie-Tooth neuropathy type 2A. Nat. Genet. 2004, 36, 449–451. [Google Scholar] [CrossRef] [PubMed]
  106. Sandoval, H.; Yao, C.K.; Chen, K.; Jaiswal, M.; Donti, T.; Lin, Y.Q.; Bayat, V.; Xiong, B.; Zhang, K.; David, G.; et al. Mitochondrial fusion but not fission regulates larval growth and synaptic development through steroid hormone production. eLife 2014, 3, e03558. [Google Scholar] [CrossRef] [PubMed]
  107. Del Amo, V.L.; Seco-Cervera, M.; García-Giménez, J.L.; Whitworth, A.J.; Pallardó, F.V.; Galindo, M.I. Mitochondrial defects and neuromuscular degeneration caused by altered expression of Drosophila Gdap1: Implications for the Charcot–Marie–Tooth neuropathy. Hum. Mol. Genet. 2014, 24, 21–36. [Google Scholar] [CrossRef] [Green Version]
  108. Storkebaum, E.; Leitão-Gonçalves, R.; Godenschwege, T.; Nangle, L.; Mejia, M.; Bosmans, I.; Ooms, T.; Jacobs, A.; Van Dijck, P.; Yang, X.L.; et al. Dominant mutations in the tyrosyl-tRNA synthetase gene recapitulate in Drosophila features of human Charcot-Marie-Tooth neuropathy. Proc. Natl. Acad. Sci. USA 2009, 106, 11782–11787. [Google Scholar] [CrossRef] [Green Version]
  109. Chihara, T.; Luginbuhl, D.; Luo, L. Cytoplasmic and mitochondrial protein translation in axonal and dendritic terminal arborization. Nat. Neurosci. 2007, 10, 828–837. [Google Scholar] [CrossRef]
  110. Duan, R.; Shi, Y.; Yu, L.; Zhang, G.; Li, J.; Lin, Y.; Guo, J.; Wang, J.; Shen, L.; Jiang, H.; et al. UBA5 Mutations Cause a New Form of Autosomal Recessive Cerebellar Ataxia. PLoS ONE 2016, 11, e0149039. [Google Scholar] [CrossRef]
  111. Chen, K.; Lin, G.; Haelterman, N.A.; Ho, T.S.-Y.; Li, T.; Li, Z.; DuRaine, L.; Graham, B.H.; Jaiswal, M.; Yamamoto, S.; et al. Loss of Frataxin induces iron toxicity, sphingolipid synthesis, and Pdk1/Mef2 activation, leading to neurodegeneration. eLife 2016, 5, e16043. [Google Scholar] [CrossRef] [Green Version]
  112. Kim, M.; Sandford, E.; Gatica, D.; Qiu, Y.; Liu, X.; Zheng, Y.; Schulman, B.A.; Xu, J.; Semple, I.; Ro, S.H.; et al. Mutation in ATG5 reduces autophagy and leads to ataxia with developmental delay. eLife 2016, 5, e12245. [Google Scholar] [CrossRef] [Green Version]
  113. Li, C.; Brazill, J.M.; Liu, S.; Bello, C.; Zhu, Y.; Morimoto, M.; Cascio, L.; Pauly, R.; Diaz-Perez, Z.; Malicdan, M.C.V.; et al. Spermine synthase deficiency causes lysosomal dysfunction and oxidative stress in models of Snyder-Robinson syndrome. Nat. Commun. 2017, 8, 1257. [Google Scholar] [CrossRef] [PubMed]
  114. Leiserson, W.M.; Forbush, B.; Keshishian, H. Drosophila glia use a conserved cotransporter mechanism to regulate extracellular volume. Glia 2010, 59, 320–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Wang, Y.; Moussian, B.; Schaeffeler, E.; Schwab, M.; Nies, A.T. The fruit fly Drosophila melanogaster as an innovative preclinical ADME model for solute carrier membrane transporters, with consequences for pharmacology and drug therapy. Drug Discov. Today 2018, 23, 1746–1760. [Google Scholar] [CrossRef] [PubMed]
  116. Yamamoto, S.; Jaiswal, M.; Charng, W.L.; Gambin, T.; Karaca, E.; Mirzaa, G.; Wiszniewski, W.; Sandoval, H.; Haelterman, N.A.; Xiong, B.; et al. A Drosophila genetic resource of mutants to study mechanisms underlying human genetic diseases. Cell 2014, 159, 200–214. [Google Scholar] [CrossRef] [Green Version]
  117. Link, N.; Chung, H.; Jolly, A.; Withers, M.; Tepe, B.; Arenkiel, B.R.; Shah, P.S.; Krogan, N.J.; Aydin, H.; Geckinli, B.B.; et al. Mutations in ANKLE2, a ZIKA Virus Target, Disrupt an Asymmetric Cell Division Pathway in Drosophila Neuroblasts to Cause Microcephaly. Dev. Cell 2019, 51, 713–729.e6. [Google Scholar] [CrossRef]
  118. Curtin, K.D.; Meinertzhagen, I.A.; Wyman, R.J. Basigin (EMMPRIN/CD147) interacts with integrin to affect cellular architecture. J. Cell Sci. 2005, 118, 2649–2660. [Google Scholar] [CrossRef] [Green Version]
  119. Huang, Y.; Mao, X.; van Jaarsveld, R.; Shu, L.; Terhal, P.A.; Jia, Z.; Xi, H.; Peng, Y.; Yan, H.; Yuan, S.; et al. Variants in CAPZA2, a member of an F-actin capping complex, cause intellectual disability and developmental delay. Hum. Mol. Genet. 2020, 29, 1537–1546. [Google Scholar] [CrossRef]
  120. Kelly, S.M.; Leung, S.W.; Pak, C.; Banerjee, A.; Moberg, K.H.; Corbett, A.H. A conserved role for the zinc finger polyadenosine RNA binding protein, ZC3H14, in control of poly(A) tail length. RNA 2014, 20, 681–688. [Google Scholar] [CrossRef] [Green Version]
  121. Nahm, M.; Lee, M.-J.; Parkinson, W.; Lee, M.; Kim, H.; Kim, Y.-J.; Kim, S.; Cho, Y.S.; Min, B.-M.; Bae, Y.C.; et al. Spartin regulates synaptic growth and neuronal survival by inhibiting BMP-mediated microtubule stabilization. Neuron 2013, 77, 680–695. [Google Scholar] [CrossRef] [Green Version]
  122. Kim, S.; Kim, J.; Park, S.; Park, J.J.; Lee, S. Drosophila Graf regulates mushroom body β-axon extension and olfactory long-term memory. Mol. Brain 2021, 14, 73. [Google Scholar] [CrossRef]
  123. Malik, B.R.; Gillespie, J.M.; Hodge, J.J. CASK and CaMKII function in the mushroom body α’/β’ neurons during Drosophila memory formation. Front. Neural Circuits 2013, 7, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Liu, Z.; Huang, Y.; Hu, W.; Huang, S.; Wang, Q.; Han, J.; Zhang, Y.Q. dAcsl, the Drosophila ortholog of acyl-CoA synthetase long-chain family member 3 and 4, inhibits synapse growth by attenuating bone morphogenetic protein signaling via endocytic recycling. J. Neurosci. 2014, 34, 2785–2796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Li, D.; Wang, Q.; Gong, N.N.; Kurolap, A.; Feldman, H.B.; Boy, N.; Brugger, M.; Grand, K.; McWalter, K.; Sacoto, M.J.; et al. Pathogenic variants in SMARCA5, a chromatin remodeler, cause a range of syndromic neurodevelopmental features. Sci. Adv. 2021, 7, eabf2066. [Google Scholar] [CrossRef] [PubMed]
  126. Chao, H.T.; Davids, M.; Burke, E.; Pappas, J.G.; Rosenfeld, J.A.; McCarty, A.J.; Davis, T.; Wolfe, L.; Toro, C.; Tifft, C.; et al. A Syndromic Neurodevelopmental Disorder Caused by De Novo Variants in EBF3. Am. J. Hum. Genet. 2017, 100, 128–137. [Google Scholar] [CrossRef] [Green Version]
  127. Farhan, S.; Nixon, K.; Everest, M.; Edwards, T.N.; Long, S.; Segal, D.; Knip, M.J.; Arts, H.H.; Chakrabarti, R.; Wang, J.; et al. Identification of a novel synaptic protein, TMTC3, involved in periventricular nodular heterotopia with intellectual disability and epilepsy. Hum. Mol. Genet. 2017, 26, 4278–4289. [Google Scholar] [CrossRef]
  128. Ansar, M.; Chung, H.L.; Al-Otaibi, A.; Elagabani, M.N.; Ravenscroft, T.A.; Paracha, S.A.; Scholz, R.; Abdel Magid, T.; Sarwar, M.T.; Shah, S.F.; et al. Bi-allelic Variants in IQSEC1 Cause Intellectual Disability, Developmental Delay, and Short Stature. Am. J. Hum. Genet. 2019, 105, 907–920. [Google Scholar] [CrossRef]
  129. Yap, Z.Y.; Strucinska, K.; Matsuzaki, S.; Lee, S.; Si, Y.; Humphries, K.; Tarnopolsky, M.A.; Yoon, W.H. A biallelic pathogenic variant in the OGDH gene results in a neurological disorder with features of a mitochondrial disease. J. Inherit. Metab. Dis. 2020, 44, 388–400. [Google Scholar] [CrossRef]
  130. Chao, Y.-H.; Robak, L.A.; Xia, F.; Koenig, M.K.; Adesina, A.; Bacino, C.A.; Scaglia, F.; Bellen, H.; Wangler, M.F. Missense variants in the middle domain of DNM1L in cases of infantile encephalopathy alter peroxisomes and mitochondria when assayed in Drosophila. Hum. Mol. Genet. 2016, 25, 1846–1856. [Google Scholar] [CrossRef] [Green Version]
  131. Shao, L.; Shuai, Y.; Wang, J.; Feng, S.; Lu, B.; Li, Z.; Zhao, Y.; Wang, L.; Zhong, Y. Schizophrenia susceptibility gene dysbindin regulates glutamatergic and dopaminergic functions via distinctive mechanisms in Drosophila. Proc. Natl. Acad. Sci. USA 2011, 108, 18831–18836. [Google Scholar] [CrossRef] [Green Version]
  132. Tamberg, L.; Sepp, M.; Timmusk, T.; Palgi, M. Introducing Pitt-Hopkins syndrome-associated mutations of TCF4 to Drosophila daughterless. Biol. Open 2015, 4, 1762–1771. [Google Scholar] [CrossRef] [Green Version]
  133. Gavilan, H.S.; Kulikauskas, R.M.; Gutmann, D.H.; Fehon, R.G. In vivo functional analysis of the human NF2 tumor suppressor gene in Drosophila. PLoS ONE 2014, 9, e90853. [Google Scholar] [CrossRef] [PubMed]
  134. Bodmer, R. Heart development in Drosophila and its relationship to vertebrates. Trends Cardiovasc. Med. 1995, 5, 21–28. [Google Scholar] [CrossRef]
  135. Bodmer, R.; Venkatesh, T.V. Heart development in Drosophila and vertebrates: Conservation of molecular mechanisms. Dev. Genet. 1998, 22, 181–186. [Google Scholar] [CrossRef]
  136. Ahmad, S.M. Conserved signaling mechanisms in Drosophila heart development. Dev. Dyn. 2017, 246, 641–656. [Google Scholar] [CrossRef] [Green Version]
  137. Souidi, A.; Jagla, K. Drosophila Heart as a Model for Cardiac Development and Diseases. Cells 2021, 10, 3078. Available online: https://www.mdpi.com/2073-4409/10/11/3078 (accessed on 13 January 2022). [CrossRef] [PubMed]
  138. Jay, P.Y.; Harris, B.S.; Maguire, C.T.; Buerger, A.; Wakimoto, H.; Tanaka, M.; Kupershmidt, S.; Roden, D.M.; Schultheiss, T.M.; O’Brien, T.X.; et al. Nkx2-5 mutation causes anatomic hypoplasia of the cardiac conduction system. J. Clin. Investig. 2004, 113, 1130–1137. [Google Scholar] [CrossRef] [Green Version]
  139. Moskowitz, I.P.; Kim, J.B.; Moore, M.L.; Wolf, C.M.; Peterson, M.A.; Shendure, J.; Nobrega, M.A.; Yokota, Y.; Berul, C.; Izumo, S.; et al. A Molecular Pathway Including Id2, Tbx5, and Nkx2-5 Required for Cardiac Conduction System Development. Cell 2007, 129, 1365–1376. [Google Scholar] [CrossRef] [Green Version]
  140. Qian, L.; Mohapatra, B.; Akasaka, T.; Liu, J.; Ocorr, K.; Towbin, J.A.; Bodmer, R. Transcription factor neuromancer/TBX20 is required for cardiac function in Drosophila with implications for human heart disease. Proc. Natl. Acad. Sci. USA 2008, 105, 19833–19838. [Google Scholar] [CrossRef] [Green Version]
  141. Qian, L.; Bodmer, R. Partial loss of GATA factor Pannier impairs adult heart function in Drosophila. Hum. Mol. Genet. 2009, 18, 3153–3163. [Google Scholar] [CrossRef] [Green Version]
  142. Qian, L.; Wythe, J.D.; Liu, J.; Cartry, J.; Vogler, G.; Mohapatra, B.; Otway, R.T.; Huang, Y.; King, I.N.; Maillet, M.; et al. Tinman/Nkx2-5 acts via miR-1 and upstream of Cdc42 to regulate heart function across species. J. Cell Biol. 2011, 193, 1181–1196. [Google Scholar] [CrossRef] [Green Version]
  143. Taghli-Lamallem, O.; Auxerre-Plantié, E.; Jagla, K. Drosophila in the heart of understanding cardiac diseases: Modeling channelopathies and cardiomyopathies in the fruitfly. J. Cardiovasc. Dev. Dis. 2016, 3, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Reiter, L.T.; Potocki, L.; Chien, S.; Gribskov, M.; Bier, E. A systematic analysis of human disease-associated gene sequences in Drosophila melanogaster. Genome Res. 2001, 11, 1114–1125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Zimmerman, M.S.; Smith, A.G.C.; Sable, C.A.; Echko, M.M.; Wilner, L.B.; Olsen, H.E.; Atalay, H.T.; Awasthi, A.; Bhutta, Z.A.; Boucher, J.L.; et al. Global, regional, and national burden of congenital heart disease, 1990–2017: A systematic analysis for the Global Burden of Disease Study 2017. Lancet Child Adolesc. Health 2020, 4, 185–200. [Google Scholar] [CrossRef] [Green Version]
  146. Pediatric Cardiac Genomics, C.; Gelb, B.; Brueckner, M.; Chung, W.; Goldmuntz, E.; Kaltman, J.; Kaski, J.P.; Kim, R.; Kline, J.; Mercer-Rosa, L.; et al. The congenital heart disease genetic network study: Rationale, design, and early results. Circ Res. 2013, 112, 698–706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Zhu, J.Y.; Fu, Y.; Nettleton, M.; Richman, A.; Han, Z. High throughput in vivo functional validation of candidate congenital heart disease genes in Drosophila. eLife 2017, 6, e22617. [Google Scholar] [CrossRef]
  148. McKenna, W.J.; Maron, B.J.; Thiene, G. Classification, epidemiology, and global burden of cardiomyopathies. Circ. Res. 2017, 121, 722–730. [Google Scholar] [CrossRef] [Green Version]
  149. Hershberger, R.E.; Hedges, D.J.; Morales, A. Dilated cardiomyopathy: The complexity of a diverse genetic architecture. Nat. Rev. Cardiol. 2013, 10, 531–547. [Google Scholar] [CrossRef]
  150. Jordan, E.; Peterson, L.; Ai, T.; Asatryan, B.; Bronicki, L.; Brown, E.; Celeghin, R.; Edwards, M.; Fan, J.; Ingles, J.; et al. Evidence-Based Assessment of Genes in Dilated Cardiomyopathy. Circulation 2021, 144, 7–19. [Google Scholar] [CrossRef]
  151. Semsarian, C.; Ingles, J.; Maron, M.S.; Maron, B.J. New perspectives on the prevalence of hypertrophic cardiomyopathy. J. Am. Coll. Cardiol. 2015, 65, 1249–1254. [Google Scholar] [CrossRef] [Green Version]
  152. Geske, J.B.; Ommen, S.R.; Gersh, B.J. Hypertrophic Cardiomyopathy: Clinical Update. JACC Heart Fail. 2018, 6, 364–375. [Google Scholar] [CrossRef]
  153. Ingles, J.; Goldstein, J.; Thaxton, C.; Caleshu, C.; Corty, E.W.; Crowley, S.B.; Dougherty, K.; Harrison, S.M.; McGlaughon, J.; Milko, L.V.; et al. Evaluating the Clinical Validity of Hypertrophic Cardiomyopathy Genes. Circ. Genom. Precis. Med. 2019, 12, e002460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Musunuru, K.; Hershberger, R.E.; Day, S.M.; Klinedinst, N.J.; Landstrom, A.P.; Parikh, V.N.; Prakash, S.; Semsarian, C.; Sturm, A.C.; American Heart Association Council on Genomic and Precision Medicine; et al. Genetic testing for inherited cardiovascular diseases: A scientific statement from the american heart association. Circ. Genom. Precis. Med. 2020, 13, e000067. [Google Scholar] [CrossRef] [PubMed]
  155. Manivannan, S.N.; Darouich, S.; Masmoudi, A.; Gordon, D.; Zender, G.; Han, Z.; Fitzgerald-Butt, S.; White, P.; McBride, K.L.; Kharrat, M.; et al. Novel frameshift variant in MYL2 reveals molecular differences between dominant and recessive forms of hypertrophic cardiomyopathy. PLoS Genet. 2020, 16, e1008639. [Google Scholar] [CrossRef]
  156. Moore, J.R.; Dickinson, M.H.; Vigoreaux, J.O.; Maughan, D.W. The effect of removing the N-terminal extension of the Drosophila myosin regulatory light chain upon flight ability and the contractile dynamics of indirect flight muscle. Biophys. J. 2000, 78, 1431–1440. [Google Scholar] [CrossRef] [Green Version]
  157. Campuzano, V.; Montermini, L.; Molto, M.D.; Pianese, L.; Cossee, M.; Cavalcanti, F.; Monros, E.; Rodius, F.; Duclos, F.; Monticelli, A.; et al. Friedreich’s ataxia: Autosomal recessive disease caused by an intronic GAA triplet repeat expansion. Science 1996, 271, 1423–1427. [Google Scholar] [CrossRef]
  158. Weidemann, F.; Liu, D.; Hu, K.; Florescu, C.; Niemann, M.; Herrmann, S.; Kramer, B.; Klebe, S.; Doppler, K.; Uceyler, N.; et al. The cardiomyopathy in Friedreich’s ataxia—New biomarker for staging cardiac involvement. Int. J. Cardiol. 2015, 194, 50–57. [Google Scholar] [CrossRef] [Green Version]
  159. Tricoire, H.; Palandri, A.; Bourdais, A.; Camadro, J.-M.; Monnier, V. Methylene blue rescues heart defects in a Drosophila model of friedreich’s ataxia. Hum. Mol. Genet. 2014, 23, 968–979. [Google Scholar] [CrossRef] [Green Version]
  160. Gonçalves, S.; Patat, J.; Guida, M.C.; Lachaussee, N.; Arrondel, C.; Helmstädter, M.; Boyer, O.; Gribouval, O.; Gubler, M.C.; Mollet, G.; et al. A homozygous KAT2B variant modulates the clinical phenotype of ADD3 deficiency in humans and flies. PLoS Genet. 2018, 14, e1007386. [Google Scholar] [CrossRef] [PubMed]
  161. Carré, C.; Szymczak, D.; Pidoux, J.; Antoniewski, C. The histone H3 acetylase dGcn5 is a key player in Drosophila melanogaster metamorphosis. Mol. Cell. Biol. 2005, 25, 8228–8238. [Google Scholar] [CrossRef] [Green Version]
  162. Casas-Tintó, S.; Arnés, M.; Ferrús, A. Drosophila enhancer-Gal4 lines show ectopic expression during development. R. Soc. Open Sci. 2017, 4, 170039. [Google Scholar] [CrossRef] [Green Version]
  163. Ocorr, K.; Reeves, N.L.; Wessells, R.J.; Fink, M.; Chen, H.S.; Akasaka, T.; Yasuda, S.; Metzger, J.M.; Giles, W.; Posakony, J.W.; et al. KCNQ potassium channel mutations cause cardiac arrhythmias in Drosophila that mimic the effects of aging. Proc. Natl. Acad. Sci. USA 2007, 104, 3943–3948. [Google Scholar] [CrossRef] [Green Version]
  164. Zhang, D.; Ke, L.; Mackovicova, K.; Van Der Want, J.J.; Sibon, O.C.; Tanguay, R.M.; Morrow, G.; Henning, R.H.; Kampinga, H.H.; Brundel, B.J. Effects of different small HSPB members on contractile dysfunction and structural changes in a Drosophila melanogaster model for Atrial Fibrillation. J. Mol. Cell. Cardiol. 2011, 51, 381–389. [Google Scholar] [CrossRef] [PubMed]
  165. Santalla, M.; Valverde, C.A.; Harnichar, E.; Lacunza, E.; Aguilar-Fuentes, J.; Mattiazzi, A.; Ferrero, P. Aging and CaMKII alter intracellular Ca2+ transients and heart rhythm in Drosophila melanogaster. PLoS ONE. 2014, 9, e101871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Limpitikul, W.B.; Viswanathan, M.C.; O’Rourke, B.; Yue, D.T.; Cammarato, A. Conservation of cardiac L-type Ca2+ channels and their regulation in Drosophila: A novel genetically-pliable channelopathic model. J. Mol. Cell. Cardiol. 2018, 119, 64–74. [Google Scholar] [CrossRef] [PubMed]
  167. Pineda, S.; Nikolova-Krstevski, V.; Leimena, C.; Atkinson, A.J.; Altekoester, A.K.; Cox, C.D.; Jacoby, A.; Huttner, I.G.; Ju, Y.K.; Soka, M.; et al. Conserved Role of the Large Conductance Calcium-Activated Potassium Channel, KCa1.1, in Sinus Node Function and Arrhythmia Risk. Circ. Genom. Precis Med. 2021, 14, e003144. [Google Scholar] [CrossRef]
  168. Yu, L.; Daniels, J.; Glaser, A.E.; Wolf, M.J. Raf-mediated cardiac hypertrophy in adult Drosophila. DMM Dis. Model. Mech. 2013, 6, 964–976. [Google Scholar]
  169. Yu, L.; Daniels, J.P.; Wu, H.; Wolf, M.J. Cardiac hypertrophy induced by active Raf depends on Yorkie-mediated transcription. Sci. Signal. 2015, 8, ra13. [Google Scholar] [CrossRef] [Green Version]
  170. Migunova, E.; Theophilopoulos, J.; Mercadante, M.; Men, J.; Zhou, C.; Dubrovsky, E.B. ELAC2/RNaseZ-linked cardiac hypertrophy in Drosophila melanogaster. DMM Dis. Model. Mech. 2021, 14, dmm048931. [Google Scholar] [CrossRef]
  171. Bloemink, M.J.; Melkani, G.C.; Dambacher, C.M.; Bernstein, S.I.; Geeves, M.A. Two Drosophila myosin transducer mutants with distinct cardiomyopathies have divergent ADP and actin affinities. J. Biol. Chem. 2011, 286, 28435–28443. [Google Scholar] [CrossRef] [Green Version]
  172. Achal, M.; Trujillo, A.S.; Melkani, G.C.; Farman, G.P.; Ocorr, K.; Viswanathan, M.C.; Kaushik, G.; Newhard, C.S.; Glasheen, B.M.; Melkani, A.; et al. A Restrictive Cardiomyopathy Mutation in an Invariant Proline at the Myosin Head/Rod Junction Enhances Head Flexibility and Function, Yielding Muscle Defects in Drosophila. J. Mol. Biol. 2016, 428, 2446–2461. [Google Scholar] [CrossRef] [Green Version]
  173. Taghli-Lamallem, O.; Akasaka, T.; Hogg, G.; Nudel, U.; Yaffe, D.; Chamberlain, J.S.; Ocorr, K.; Bodmer, R. Dystrophin deficiency in Drosophila reduces lifespan and causes a dilated cardiomyopathy phenotype. Aging Cell 2008, 7, 237–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Kim, I.M.; Wolf, M.J. Serial examination of an inducible and reversible dilated cardiomyopathy in individual adult Drosophila. PLoS ONE 2009, 4, e7132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Viswanathan, M.C.; Kaushik, G.; Engler, A.J.; Lehman, W.; Cammarato, A. A Drosophila melanogaster model of diastolic dysfunction and cardiomyopathy based on impaired troponin-T function. Circ Res. 2014, 114, e6–e17. [Google Scholar]
  176. Selma-Soriano, E.; Casillas-Serra, C.; Artero, R.; Llamusi, B.; Navarro, J.A.; Redón, J. Rabphilin silencing causes dilated cardiomyopathy in a Drosophila model of nephrocyte damage. Sci. Rep. 2021, 11, 15287. [Google Scholar] [CrossRef]
  177. Allikian, M.J.; Bhabha, G.; Dospoy, P.; Heydemann, A.; Ryder, P.; Earley, J.U.; Wolf, M.J.; Rockman, H.A.; McNally, E.M. Reduced life span with heart and muscle dysfunction in Drosophila sarcoglycan mutants. Hum. Mol. Genet. 2007, 16, 2933–2943. [Google Scholar] [CrossRef] [Green Version]
  178. Taghli-Lamallem, O.; Jagla, K.; Chamberlain, J.S.; Bodmer, R. Mechanical and non-mechanical functions of Dystrophin can prevent cardiac abnormalities in Drosophila. Exp. Gerontol. 2014, 49, 26–34. [Google Scholar] [CrossRef] [Green Version]
  179. Gao, Q.Q.; Wyatt, E.; Goldstein, J.A.; LoPresti, P.; Castillo, L.M.; Gazda, A.; Petrossian, N.; Earley, J.U.; Hadhazy, M.; Barefield, D.Y.; et al. Reengineering a transmembrane protein to treat muscular dystrophy using exon skipping. J. Clin. Investig. 2015, 125, 4186–4195. [Google Scholar] [CrossRef] [Green Version]
  180. Tang, M.; Yuan, W.; Fan, X.; Liu, M.; Bodmer, R.; Ocorr, K.; Wu, X. Pygopus maintains heart function in aging Drosophila independently of canonical Wnt signaling. Circ. Cardiovasc. Genet. 2013, 6, 472–480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Tang, M.; Yuan, W.; Bodmer, R.; Wu, X.; Ocorr, K. The role of pygopus in the differentiation of intracardiac valves in Drosophila. Genesis 2014, 52, 19–28. [Google Scholar] [CrossRef] [Green Version]
  182. Kramps, T.; Peter, O.; Brunner, E.; Nellen, D.; Froesch, B.; Chatterjee, S.; Murone, M.; Zullig, S.; Basler, K. Wnt/wingless signaling requires BCL9/legless-mediated recruitment of pygopus to the nuclear beta-catenin-TCF complex. Cell 2002, 109, 47–60. [Google Scholar] [CrossRef] [Green Version]
  183. Thomson, K.L.; Ormondroyd, E.; Harper, A.R.; Dent, T.; McGuire, K.; Baksi, J.; Blair, E.; Brennan, P.; Buchan, R.; Bueser, T.; et al. Analysis of 51 proposed hypertrophic cardiomyopathy genes from genome sequencing data in sarcomere negative cases has negligible diagnostic yield. Genet. Med. 2019, 21, 1576–1584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Micheu, M.M.; Popa-Fotea, N.M.; Oprescu, N.; Dorobantu, M.; Ratiu, A.C.; Ecovoiu, A.A. NGS data validated by Sanger sequencing reveal a puzzling small deletion of MYBPC3 gene associated with hypertrophic cardiomyopathy. Rom. Biotechnol. Lett. 2019, 24, 91–99. [Google Scholar] [CrossRef]
  185. Micheu, M.M.; Popa-Fotea, N.M.; Oprescu, N.; Bogdan, S.; Dan, M.; Deaconu, A.; Dorobantu, L.; Gheorghe-Fronea, O.; Greavu, M.; Iorgulescu, C.; et al. Yield of Rare Variants Detected by Targeted Next-Generation Sequencing in a Cohort of Romanian Index Patients with Hypertrophic Cardiomyopathy. Diagnostics 2020, 10, 1061. [Google Scholar] [CrossRef] [PubMed]
  186. Alimohamed, M.Z.; Johansson, L.F.; Posafalvi, A.; Boven, L.G.; van Dijk, K.K.; Walters, L.; Vos, Y.J.; Westers, H.; Hoedemaekers, Y.M.; Sinke, R.J.; et al. Diagnostic yield of targeted next generation sequencing in 2002 Dutch cardiomyopathy patients. Int. J. Cardiol. 2021, 332, 90–104. [Google Scholar] [CrossRef]
  187. Kim, A.R.; Choi, K.W. TRiC/CCT chaperonins are essential for organ growth by interacting with insulin/TOR signaling in Drosophila. Oncogene 2019, 38, 4739–4754. [Google Scholar] [CrossRef] [Green Version]
  188. Jung, W.H.; Liu, C.C.; Yu, Y.L.; Chang, Y.C.; Lien, W.Y.; Chao, H.C.; Huang, S.Y.; Kuo, C.H.; Ho, H.C.; Chan, C.C. Lipophagy prevents activity-dependent neurodegeneration due to dihydroceramide accumulation in vivo. EMBO Rep. 2017, 18, 1150–1165. [Google Scholar] [CrossRef] [PubMed]
  189. Muyrers-Chen, I.; Rozovskaia, T.; Lee, N.; Kersey, J.H.; Nakamura, T.; Canaani, E.; Paro, R. Expression of leukemic MLL fusion proteins in Drosophila affects cell cycle control and chromosome morphology. Oncogene 2004, 23, 8639–8648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Perkins, L.A.; Johnson, M.R.; Melnick, M.B.; Perrimon, N. The nonreceptor protein tyrosine phosphatase corkscrew functions in multiple receptor tyrosine kinase pathways in Drosophila. Dev. Biol. 1996, 180, 63–81. [Google Scholar] [CrossRef] [Green Version]
  191. Tan, K.L.; Haelterman, N.A.; Kwartler, C.S.; Regalado, E.S.; Lee, P.T.; Nagarkar-Jaiswal, S.; Guo, D.C.; Duraine, L.; Wangler, M.F.; University of Washington Center for Mendelian Genomics; et al. Ari-1 regulates myonuclear organization together with parkin and is associated with aortic aneurysms. Dev. Cell 2018, 45, 226–244. [Google Scholar] [CrossRef] [Green Version]
  192. Mirzoyan, Z.; Sollazzo, M.; Allocca, M.; Valenza, A.M.; Grifoni, D.; Bellosta, P. Drosophila melanogaster: A Model Organism to Study Cancer. Front. Genet. 2019, 10, 51. [Google Scholar] [CrossRef] [Green Version]
  193. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Kumar, M.; Lechel, A.; Gunes, C. Telomerase: The devil inside. Genes 2016, 7, 43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Vander Heiden, M.G.; DeBerardinis, R.J. Understanding the intersections between metabolism and cancer biology. Cell 2017, 168, 657–669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Millburn, G.H.; Crosby, M.A.; Gramates, L.S.; Tweedie, S.; FlyBase, C. FlyBase portals to human disease re-search using Drosophila models. Dis. Model. Mech. 2016, 9, 245–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Grifoni, D.; Garoia, F.; Schimanski, C.C.; Schmitz, G.; Laurenti, E.; Galle, P.R.; Pession, A.; Cavicchi, S.; Strand, D. The human protein Hugl-1 substitutes for Drosophila Lethal giant larvae tumour suppressor function in vivo. Oncogene 2004, 23, 8688–8694. [Google Scholar] [CrossRef] [Green Version]
  198. Tao, W.; Zhang, S.; Turenchalk, G.S.; Stewart, R.A.; St John, M.A.; Chen, W.; Xu, T. Human homologue of the Drosophila melanogaster lats tumour suppressor modulates CDC2 activity. Nat. Genet. 1999, 21, 177–181. [Google Scholar] [CrossRef]
  199. Dow, L.E.; Brumby, A.M.; Muratore, R.; Coombe, M.L.; Sedelies, K.A.; Trapani, J.A.; Russell, S.M.; Richardson, H.E.; Humbert, P.O. hScrib is a functional homologue of the Drosophila tumour suppressor Scribble. Oncogene 2003, 22, 9225–9230. [Google Scholar] [CrossRef] [Green Version]
  200. Benchabane, H.; Xin, N.; Tian, A.; Hafler, B.P.; Nguyen, K.; Ahmed, A.; Ahmed, Y. Jerky/Earthbound facilitates cell-specific Wnt/Wingless signalling by modulating β-catenin-TCF activity. EMBO J. 2011, 30, 1444–1458. [Google Scholar] [CrossRef] [Green Version]
  201. Drusenheimer, N.; Migdal, B.; Jäckel, S.; Tveriakhina, L.; Scheider, K.; Schulz, K.; Gröper, J.; Köhrer, K.; Klein, T. The Mam-malian Orthologs of Drosophila Lgd, CC2D1A and CC2D1B, Function in the Endocytic Pathway, but Their Individual Loss of Function Does Not Affect Notch Signalling. PLoS Genet. 2015, 11, e1005749. [Google Scholar] [CrossRef] [Green Version]
  202. D’Brot, A.; Kurtz, P.; Regan, E.; Jakubowski, B.; Abrams, J.M. A platform for interrogating cancer-associated p53 alleles. Oncogene 2016, 36, 286–291. [Google Scholar] [CrossRef] [Green Version]
  203. Bras, S.; Martin-Lanneree, S.; Gobert, V.; Auge, B.; Breig, O.; Sanial, M.; Yamaguchi, M.; Haenlin, M.; Plessis, A.; Waltzer, L. Myeloid leukemia factor is a conserved regulator of RUNX transcription factor activity involved in hematopoiesis. Proc. Natl. Acad. Sci. USA 2012, 109, 4986–4991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Ludlow, C.; Choy, R.; Blochlinger, K. Functional analysis of Drosophila and mammalian cut proteins in files. Dev. Biol. 1996, 178, 149–159. [Google Scholar] [CrossRef] [Green Version]
  205. Brumby, A.M.; Richardson, H.E. scribble mutants cooperate with oncogenic Ras or Notch to cause neo-plastic overgrowth in Drosophila. EMBO J. 2003, 22, 5769–5779. [Google Scholar] [CrossRef] [PubMed]
  206. Lee, T.; Luo, L. Mosaic analysis with a repressible cell marker for studies of gene function in neuronal morphogenesis. Neuron 1999, 22, 451–461. [Google Scholar] [CrossRef] [Green Version]
  207. Pagliarini, R.A.; Xu, T. A genetic screen in Drosophila for metastatic behavior. Science 2003, 302, 1227–1231. [Google Scholar] [CrossRef] [PubMed]
  208. Tipping, M.; Perrimon, N. Drosophila as a model for context-dependent tumorigenesis. J. Cell. Physiol. 2014, 229, 27–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Wu, Y.; Zhou, B.P. Inflammation: A driving force speeds cancer metastasis. Cell Cycle 2009, 8, 3267–3273. [Google Scholar] [CrossRef] [Green Version]
  210. Grzeschik, N.A.; Parsons, L.M.; Richardson, H.E. Lgl, the SWH pathway and tumorigenesis: It’s a matter of context & competition! Cell Cycle 2010, 9, 3222–3232. [Google Scholar] [CrossRef] [Green Version]
  211. Thiery, J.P. Epithelial-mesenchymal transitions in tumour progression. Nat. Rev. Cancer 2002, 2, 442–454. [Google Scholar] [CrossRef]
  212. Woodhouse, E.; Hersperger, E.; Shearn, A. Growth, metastasis, and invasiveness of Drosophila tumors caused by mutations in specific tumor suppressor genes. Dev. Genes Evol. 1998, 207, 542–550. [Google Scholar] [CrossRef]
  213. Woodhouse, E.; Hersperger, E.; Stetler-Stevenson, W.G.; Liotta, L.A.; Shearn, A. Increased type IV collagenase in lgl-induced invasive tumors of Drosophila. Cell Growth Differ. 1994, 5, 151–159. [Google Scholar] [PubMed]
  214. Xu, J.; Liu, L.Z.; Deng, X.F.; Timmons, L.; Hersperger, E.; Steeg, P.S.; Veron, M.; Shearn, A. The Enzymatic Activity of Drosophila AWD/NDP Kinase Is Necessary but Not Suffi-cient for Its Biological Function. Dev. Biol. 1996, 177, 544–557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Gateff, E. Malignant neoplasms of genetic origin in Drosophila melanogaster. Science 1978, 200, 1448–1459. [Google Scholar] [CrossRef]
  216. Stoker, M.G.; Shearer, M.; O’Neill, C. Growth inhibition of polyoma-transformed cells by contact with static normal fibroblasts. J. Cell Sci. 1966, 1, 297–310. [Google Scholar] [CrossRef]
  217. Baker, N.E.; Li, W. Cell competition and its possible relation to cancer. Cancer Res. 2008, 68, 5505–5507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Moreno, E. Is cell competition relevant to cancer? Nat. Rev. Cancer 2008, 8, 141–147. [Google Scholar] [CrossRef] [PubMed]
  219. Vincent, J.-P.; Fletcher, A.G.; Baena-Lopez, L.A. Mechanisms and mechanics of cell competition in epithelia. Nat. Rev. Mol. Cell Biol. 2013, 14, 581–591. [Google Scholar] [CrossRef] [Green Version]
  220. Vincent, J.-P.; Kolahgar, G.; Gagliardi, M.; Piddini, E. Steep differences in wingless signaling trigger Myc-independent competitive cell interactions. Dev. Cell 2011, 21, 366–374. [Google Scholar] [CrossRef] [Green Version]
  221. McCartney, B.M.; Price, M.H.; Webb, R.L.; Hayden, M.A.; Holot, L.M.; Zhou, M.; Bejsovec, A.; Peifer, M. Testing hypotheses for the functions of APC family proteins using null and truncation alleles in Drosophila. Development 2006, 133, 2407–2418. [Google Scholar] [CrossRef] [Green Version]
  222. Jiang, H.; Grenley, M.O.; Bravo, M.-J.; Blumhagen, R.Z.; Edgar, B.A. EGFR/Ras/MAPK signaling mediates adult midgut epithelial homeostasis and regeneration in Drosophila. Cell Stem Cell 2011, 8, 84–95. [Google Scholar] [CrossRef] [Green Version]
  223. Jiang, H.; Edgar, B.A. Intestinal stem cell function in Drosophila and mice. Curr. Opin. Genet. Dev. 2012, 22, 354–360. [Google Scholar] [CrossRef] [Green Version]
  224. Patel, P.H.; Edgar, B.A. Tissue design: How Drosophila tumors remodel their neighborhood. Semin. Cell Dev. Biol. 2014, 28, 86–95. [Google Scholar] [CrossRef] [PubMed]
  225. Cordero, J.; Vidal, M.; Sansom, O. APC as a master regulator of intestinal homeostasis and transformation: From flies to vertebrates. Cell Cycle 2009, 8, 2926–2931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Lee, W.C.; Beebe, K.; Sudmeier, L.; Micchelli, C.A. Adenomatous polyposis coli regulates Drosophila intestinal stem cell proliferation. Development 2009, 136, 2255–2264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Yu, X.; Waltzer, L.; Bienz, M. A new Drosophila APC homologue associated with adhesive zones of epithelial cells. Nat. Cell Biol. 1999, 1, 144–151. [Google Scholar] [CrossRef]
  228. Tian, A.; Benchabane, H.; Wang, Z.; Zimmerman, C.; Xin, N.; Perochon, J.; Kalna, G.; Sansom, O.J.; Cheng, C.; Cordero, J.; et al. Intestinal stem cell overproliferation resulting from inactivation of the APC tumor suppressor requires the transcription cofactors Earthbound and Erect wing. PLoS Genet. 2017, 13, e1006870. [Google Scholar] [CrossRef] [Green Version]
  229. Barker, N.; Ridgway, R.A.; van Es, J.H.; van de Wetering, M.; Begthel, H.; van den Born, M.; Danenberg, E.; Clarke, A.R.; Sansom, O.J.; Clevers, H. Crypt stem cells as the cells-of-origin of intestinal cancer. Nature 2009, 457, 608–611. [Google Scholar] [CrossRef]
  230. Benchabane, H.; Hughes, E.G.; Takacs, C.M.; Baird, J.R.; Ahmed, Y. Adenomatous polyposis coli is present near the minimal level required for accurate graded responses to the Wingless morphogen. Development 2008, 135, 963–971. [Google Scholar] [CrossRef] [Green Version]
  231. Ahmed, Y.; Hayashi, S.; Levine, A.; Wieschaus, E. Regulation of armadillo by a Drosophila APC inhibits neuronal apoptosis during retinal development. Cell 1998, 93, 1171–1182. [Google Scholar] [CrossRef] [Green Version]
  232. Xin, N.; Benchabane, H.; Tian, A.; Nguyen, K.; Klofas, L.; Ahmed, Y. Erect Wing facilitates context-dependent Wnt/Wingless signaling by recruiting the cell-specific Armadillo-TCF adaptor Earthbound to chromatin. Development 2011, 138, 4955–4967. [Google Scholar] [CrossRef] [Green Version]
  233. Söderholm, S.; Cantù, C. The WNT/β-catenin dependent transcription: A tissue-specific business. WIREs Syst. Biol. Med. 2020, 13, e1511. [Google Scholar] [CrossRef] [PubMed]
  234. Chang, C.H.; Lai, L.C.; Cheng, H.C.; Chen, K.R.; Syue, Y.Z.; Lu, H.C.; Lin, W.Y.; Chen, S.H.; Huang, H.S.; Shiau, A.L.; et al. TBK1-associated protein in endolysosomes (TAPE) is an innate immune regulator modulating the TLR3 and TLR4 signaling pathways. J. Biol. Chem. 2011, 286, 7043–7051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Chen, K.R.; Chang, C.H.; Huang, C.Y.; Lin, C.Y.; Lin, W.Y.; Lo, Y.C.; Yang, C.Y.; Hsing, E.W.; Chen, L.F.; Shih, S.R.; et al. TBK1-associated protein in endolysosomes (TAPE)/CC2D1A is a key regulator linking RIG-I-like receptors to antiviral immunity. J. Biol. Chem. 2012, 287, 32216–32221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Rothenberg, E.V.; Taghon, T. Molecular genetics of T cell development. Annu. Rev. Immunol. 2005, 23, 601–649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Schweisguth, F. Regulation of notch signaling activity. Curr. Biol. 2004, 14, R129–R138. [Google Scholar] [CrossRef]
  238. Mumm, J.S.; Schroeter, E.H.; Saxena, M.T.; Griesemer, A.; Tian, X.; Pan, D.J.; Ray, W.J.; Kopan, R. A ligand-induced extracellular cleavage regulates gamma-secretase-like proteolytic activation of Notch1. Mol. Cell 2000, 5, 197–206. [Google Scholar] [CrossRef]
  239. Bryant, P.J.; Schubiger, G. Giant and duplicated imaginal discs in a new lethal mutant of Drosophila melanogaster. Dev. Biol. 1971, 24, 233–263. [Google Scholar] [CrossRef]
  240. Jaekel, R.; Klein, T. The Drosophila Notch inhibitor and tumor suppressor gene lethal (2) giant discs encodes a conserved regulator of endosomal trafficking. Dev. Cell 2006, 11, 655–669. [Google Scholar] [CrossRef]
  241. Klein, T. The tumour suppressor gene l(2)giant discs is required to restrict the activity of Notch to the dorsoventral boundary during Drosophila wing development. Dev. Biol. 2003, 255, 313–333. [Google Scholar] [CrossRef] [Green Version]
  242. Parr, C.; Watkins, G.; Jiang, W.G. The possible correlation of Notch-1 and Notch-2 with clinical outcome and tumour clinicopathological parameters in human breast cancer. Int. J. Mol. Med. 2004, 14, 779–786. [Google Scholar] [CrossRef]
  243. Jin, M.M.; Ye, Y.Z.; Qian, Z.D.; Zhang, Y.B. Notch signaling molecules as prognostic biomarkers for non-small cell lung cancer. Oncol. Lett. 2015, 10, 3252–3260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Ye, J.; Wen, J.; Ning, Y.; Li, Y. Higher notch expression implies poor survival in pancreatic ductal adenocarcinoma: A systematic review and meta-analysis. Pancreatology 2018, 18, 954–961. [Google Scholar] [CrossRef] [PubMed]
  245. Yuan, X.; Wu, H.; Xu, H.; Han, N.; Chu, Q.; Yu, S.; Chen, Y.; Wu, K. Meta-analysis reveals the correlation of Notch signaling with non-small cell lung cancer progression and prognosis. Sci. Rep. 2015, 5, 10338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Harnish, J.M.; Link, N.; Yamamoto, S. Drosophila as a Model for Infectious Diseases. Int. J. Mol. Sci. 2021, 22, 2724. [Google Scholar] [CrossRef]
  247. Evans, A.S. Limitations of the Henle—Koch postulates. In Causation and Disease; Springer: Boston, MA, USA, 1993. [Google Scholar] [CrossRef]
  248. Florescu, S.A.; Cotar, A.I.; Popescu, C.P.; Ceianu, C.S.; Zaharia, M.; Vancea, G.; Codreanu, D.; Badescu, D.; Ceausu, E. First Two Imported Cases of Zika Virus Infections in Romania. Vector Borne Zoonotic Dis. 2017, 17, 354–357. [Google Scholar] [CrossRef] [PubMed]
  249. Oehler, E.; Watrin, L.; Larre, P.; Leparc-Goffart, I.; Lastère, S.; Valour, F.; Baudouin, L.; Mallet, H.P.; Musso, D.; Ghawche, F. Zika Virus Infection Complicated by Guillain-Barré Syndrome Acase Report, French Polynesia, December 2013. Eurosurveillance 2014, 19, 20720. [Google Scholar] [CrossRef] [Green Version]
  250. Mlakar, J.; Korva, M.; Tul, N.; Popović, M.; Poljšak-Prijatelj, M.; Mraz, J.; Kolenc, M.; Resman Rus, K.; Vesnaver Vipotnik, T.; Fabjan Vodušek, V.; et al. Zika Virus Associated with Microcephaly. N. Engl. J. Med. 2016, 374, 951–958. [Google Scholar] [CrossRef]
  251. Klaitong, P.; Smith, D.R. Roles of non-structural protein 4A in flavivirus infection. Viruses 2021, 13, 2077. [Google Scholar] [CrossRef]
  252. Shah, P.S.; Link, N.; Jang, G.M.; Sharp, P.P.; Zhu, T.; Swaney, D.L.; Johnson, J.R.; Von Dollen, J.; Ramage, H.R.; Sat-kamp, L.; et al. Comparative Flavivirus-host protein interaction mapping reveals mechanisms of dengue and Zika virus pathogenesis. Cell 2018, 175, 1931–1945. [Google Scholar] [CrossRef] [Green Version]
  253. Shaheen, R.; Maddirevula, S.; Ewida, N.; Alsahli, S.; Abdel-Salam, G.M.H.; Zaki, M.S.; Al Tala, S.; Alhashem, A.; Softah, A.; Al-Owain, M.; et al. Genomic and phenotypic delineation of congenital microcephaly. Genet. Med. 2018, 21, 545–552. [Google Scholar] [CrossRef]
  254. Link, N.; Bellen, H.J.; Dunwoodie, S.; Wallingford, J. Using Drosophila to drive the diagnosis and understand the mechanisms of rare human diseases. Development 2020, 147, dev191411. [Google Scholar] [CrossRef]
  255. Link, N.; Chung, H.; Jolly, A.; Withers, M.; Tepe, B.; Arenkiel, B.R.; Shah, P.S.; Krogan, N.J.; Aydin, H.; Geckinli, B.B.; et al. Ankle2, a Target of Zika Virus, Controls Asymmetric Cell Division of Neuroblasts and Uncovers a Novel Microcephaly Pathway. bioRxiv 2019, 611384. [Google Scholar] [CrossRef]
  256. Almagor, L.; Ufimtsev, I.S.; Ayer, A.; Li, J.; Weis, W.I. Structural insights into the aPKC regulatory switch mechanism of the human cell polarity protein lethal giant larvae. Proc. Natl. Acad. Sci. USA 2019, 116, 10804–10812. [Google Scholar] [CrossRef] [Green Version]
  257. Gonzaga-Jauregui, C.; Lotze, T.; Jamal, L.; Penney, S.; Campbell, I.M.; Pehlivan, D.; Hunter, J.V.; Woodbury, S.L.; Raymond, G.; Adesina, A.M.; et al. Mutations in VRK1 associated with complex motor and sensory axonal neuropathy plus microcephaly. JAMA Neurol. 2013, 70, 1491–1498. [Google Scholar] [PubMed] [Green Version]
  258. Yakulov, T.; Günesdogan, U.; Jäckle, H.; Herzig, A. Bällchen participates in proliferation control and prevents the differentiation of Drosophila melanogaster neuronal stem cells. Biol. Open 2014, 3, 881–886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Ali, M.; Nelson, A.; Lopez, A.L.; Sack, D. global burden of cholera in endemic countries. PLoS Neglected Trop. Dis. 2015, 9, e0003832. Available online: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4455997/ (accessed on 15 January 2022). [CrossRef] [PubMed] [Green Version]
  260. Israil, A.; Balotescu, C.; Damian, M.; Dinu, C.; Bucurenci, N. Comparative study of different methods for detection of toxic and other enzymatic factors in Vibrio cholerae strains. Rom. Arch. Microbiol. Immunol. 2005, 63, 63–77. [Google Scholar]
  261. Bhuin, T.; Roy, J.K. Rab11 in disease progression. Int. J. Mol. Cell. Med. 2015, 4, 1–8. [Google Scholar] [PubMed]
  262. Guichard, A.; Cruz-Moreno, B.; Aguilar, B.; van Sorge, N.; Kuang, J.; Kurkciyan, A.A.; Wang, Z.; Hang, S.; de Chambrun, G.P.P.; McCole, D.F.; et al. Cholera Toxin Disrupts Barrier Function by Inhibiting Exocyst-Mediated Trafficking of Host Proteins to Intestinal Cell Junctions. Cell Host Microbe 2013, 14, 294–305. [Google Scholar] [CrossRef] [Green Version]
  263. Ferlay, J.; Bray, P.; Parkin, D.M. Globocan 2000: Cancer Incidence, Mortality and Prevalence Worldwide, Version 1.0; IARC Cancer Base No. 5.; IARC Press: Lyon, France, 2001. [Google Scholar]
  264. Chen, Y.; Segers, S.; Blaser, M.J. Association between Helicobacter pylori and mortality in the NHANES III study. Gut 2013, 62, 1262–1269. [Google Scholar] [CrossRef] [Green Version]
  265. Ilie, M.; Dascalu, L.; Macovei, R.A. Helicobacter Pylori Cag A Antibodies and Their Clinical Implications: Correlation of Helicobacter Pylori CagA Antibodies with Treatment Resistance, Bleeding Ulcer and Gastric Cance; LAP LAMBERT Academic Publishing: Saarbrucken, Germany, 2014; ISBN1 103659526630. ISBN2 13978-3659526633. [Google Scholar]
  266. Hatakeyama, M.; Higashi, H. Helicobacter Pylori CagA: A New Paradigm for Bacterial Carcinogenesis. Cancer Sci. 2005, 96, 835–843. [Google Scholar] [CrossRef] [PubMed]
  267. Butti, R.; Das, S.; Gunasekaran, V.P.; Yadav, A.S.; Kumar, D.; Kundu, G.C. Receptor tyrosine kinases (RTKs) in breast cancer: Signaling, therapeutic implications and challenges. Mol. Cancer 2018, 17, 34. [Google Scholar] [CrossRef] [Green Version]
  268. Saadat, I.; Higashi, H.; Obuse, C.; Umeda, M.; Murata-Kamiya, N.; Saito, Y.; Lu, H.; Ohnishi, N.; Azuma, T.; Suzuki, A.; et al. Helicobacter pylori CagA targets PAR1/MARK kinase to disrupt epithelial cell polarity. Nature 2007, 447, 330–333. [Google Scholar] [CrossRef]
  269. Hatakeyama, M. Structure and function of Helicobacter pylori CagA, the first-identified bacterial protein involved in human cancer. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2017, 93, 196–219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Tateno, M.; Nishida, Y.; Adachi-Yamada, T. Regulation of JNK by Src during Drosophila Development. Science 2000, 287, 324–327. [Google Scholar] [CrossRef] [PubMed]
  271. Yong, X.; Tang, B.; Li, B.S.; Xie, R.; Hu, C.J.; Luo, G.; Qin, Y.; Dong, H.; Yang, S.M. Helicobacter pylori virulence factor CagA promotes tumorigenesis of gastric cancer via multiple signaling pathways. Cell Commun. Signal. 2015, 13, 30. [Google Scholar] [CrossRef] [Green Version]
  272. Wandler, A.M.; Guillemin, K. Transgenic expression of the Helicobacter pylori virulence factor CagA promotes apoptosis or tumorigenesis through JNK activation in Drosophila. PLOS Pathog. 2012, 8, e1002939. [Google Scholar] [CrossRef] [Green Version]
  273. Igaki, T.; Pagliarini, R.A.; Xu, T. Loss of Cell Polarity Drives Tumor Growth and Invasion through JNK Activation in Drosophila. Curr. Biol. 2006, 16, 1139–1146. [Google Scholar] [CrossRef] [Green Version]
  274. Wu, M.; Pastor-Pareja, J.C.; Xu, T. Interaction between RasV12 and Scribbled Clones Induces Tumour Growth and Invasion. Nature 2010, 463, 545–548. [Google Scholar] [CrossRef] [Green Version]
  275. D’Souza, J.; Cheah, P.Y.; Gros, P.; Chia, W.; Rodrigues, V. Functional complementation of the malvolio mutation in the taste pathway of Drosophila melanogaster by the human natural resistance-associated macrophage protein 1 (Nramp-1). J. Exp. Biol. 1999, 202, 1909–1915, Printed in Great Britain © The Company of Biologists Limited 1999 JEB1976S. [Google Scholar] [CrossRef]
  276. Cellier, M.; Belouchi, A.; Gros, P. Resistance to intracellular infections: Comparative genomic analysis of Nramp. Trends Genet. 1996, 12, 201–204. [Google Scholar] [CrossRef]
  277. Rodrigues, V.; Cheah, P.Y.K.; Chia, W. Malvolio, the Drosophila homologue of mouse NRAMP-1( Bcg), is expressed in macrophages and in the nervous system and is required for normal taste behavior. EMBO J. 1995, 14, 3007–3020. [Google Scholar] [CrossRef] [PubMed]
  278. Orgad, S.; Nelson, H.; Segal, D.; Nelson, N. Metal ions suppress the abnormal taste behavior of the Drosophila mutant malvolio. J. Exp. Biol. 1998, 201, 115–120. [Google Scholar] [CrossRef] [PubMed]
  279. Abel, L.; Sanchez, F.O.; Oberti, T.N.V.; Hoa, L.V.; Lap, V.D.; Skamene, E.; Lagrange, P.H.; Schurr, E. Susceptibility to leprosy is linked to the human NRAMP1 gene. J. Infect. Dis. 1998, 177, 133–145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Bellamy, R.; Ruwende, C.; Corrah, T.; McAdam, K.P.; Whittle, H.C.; Hill, A.V. Variations in the NRAMP1 gene and susceptibility to tuberculosis in West Africans. N. Engl. J. Med. 1998, 338, 640–644. [Google Scholar] [CrossRef]
  281. Gertler, F.B.; Comer, A.R.; Juang, J.-L.; Ahern, S.M.; Clark, M.J.; Liebl, E.C.; Hoffmann, F.M. Enabled, a dosage-sensitive suppressor of mutations in the Drosophila Abl tyrosine kinase, encodes an Abl substrate with SH3-binding properties. Genes Dev. 1995, 9, 521–533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Symons, M.; Derry, J.D.J.; Karlak, B.; Jiang, S.; Lemahieu, V.; McCormick, F.F.U.; Abo, A. Wiskott-Aldrich syndrome protein, a novel effector for the GTPase CDC42Hs, is implicated in actin polymerization. Cell 1996, 84, 723–734. [Google Scholar] [CrossRef] [Green Version]
  283. Ahern-Djamali, S.M.; Comer, A.R.; Bachmann, C.; Kastenmeier, A.S.; Reddy, S.K.; Beckerle, M.C.; Walter, U.; Hoffmann, F.M. Mutations in Drosophila enabled and rescue by human vasodilator-stimulated phosphoprotein (VASP) indicate important functional roles for Ena/VASP homology domain 1 (EVH1) and EVH2 domains. Mol. Biol. Cell 1998, 9, 2157–2171. [Google Scholar] [CrossRef] [Green Version]
  284. Muñoz-Alarcón, A.; Pavlovic, M.; Wismar, J.; Schmitt, B.; Eriksson, M.; Kylsten, P.; Dushay, M.S. Characterization of lamin mutation phenotypes in Drosophila and comparison to human laminopathies. PLoS ONE 2007, 2, e532. [Google Scholar] [CrossRef] [Green Version]
  285. Beard, G.S.; Bridger, J.M.; Kill, I.R.; Tree, D.R. Towards a Drosophila model of Hutchinson-Gilford progeria syndrome. Biochem. Soc. Trans. 2008, 36 Pt 6, 1389–1392. [Google Scholar] [CrossRef]
  286. Tsurumi, A.; Li, W.X. Aging mechanisms—A perspective mostly from Drosophila. Adv. Genet. 2020, 1, e10026. [Google Scholar] [CrossRef]
Figure 1. General outline of the heterologous rescue of D. melanogaster mutant phenotypes determined by lethal LOF alleles of dGOIs with transgenic cDNAs corresponding to the orthologous hGOIs associated with a hGD. (1) Bioinformatics analysis is deployed to search in the fruit fly genome for an orthologous gene for a human gene associated with a genetic disorder. (2) A cDNA copy of the wild-type allele of hGOI is cloned into an insertional vector under the control of a UAS enhancer sensitive to GAL4 activator. The UAS–cDNA construct is delivered by micro-injection into mutant embryos heterozygous for dGOILOF allele (obtained by targeted mutagenesis), and the resulting transgenic adults are subsequently crossed with a heterozygous dGOILOF fruit fly strain able to produce the GAL4 activator. (3) Transgenic analysis of descendant dGOILOF/dGOILOF flies that contain a functional hGOI activated by GAL4 may reveal two distinct situations. YES (complete or partial heterologous rescue). The functional hGOI1 encodes a human protein of interest (hPOI1) which is able to properly interplay with an interacting protein (dIPα) in D. melanogaster molecular background; since dIPα is the proximal interactor of the normal protein encoded by wild-type copy of dGOI1, the correct interaction between hPOI1 and dIPα rescues lethality of LOF/LOF transgenics. NO (heterologous rescue fails). On the other hand, the human protein hPOI2, encoded by a different hGOI2 transgene, does not interact accurately with dIPβ and the heterologous rescue fails. Created with BioRender.com (accessed on 23 February 2022).
Figure 1. General outline of the heterologous rescue of D. melanogaster mutant phenotypes determined by lethal LOF alleles of dGOIs with transgenic cDNAs corresponding to the orthologous hGOIs associated with a hGD. (1) Bioinformatics analysis is deployed to search in the fruit fly genome for an orthologous gene for a human gene associated with a genetic disorder. (2) A cDNA copy of the wild-type allele of hGOI is cloned into an insertional vector under the control of a UAS enhancer sensitive to GAL4 activator. The UAS–cDNA construct is delivered by micro-injection into mutant embryos heterozygous for dGOILOF allele (obtained by targeted mutagenesis), and the resulting transgenic adults are subsequently crossed with a heterozygous dGOILOF fruit fly strain able to produce the GAL4 activator. (3) Transgenic analysis of descendant dGOILOF/dGOILOF flies that contain a functional hGOI activated by GAL4 may reveal two distinct situations. YES (complete or partial heterologous rescue). The functional hGOI1 encodes a human protein of interest (hPOI1) which is able to properly interplay with an interacting protein (dIPα) in D. melanogaster molecular background; since dIPα is the proximal interactor of the normal protein encoded by wild-type copy of dGOI1, the correct interaction between hPOI1 and dIPα rescues lethality of LOF/LOF transgenics. NO (heterologous rescue fails). On the other hand, the human protein hPOI2, encoded by a different hGOI2 transgene, does not interact accurately with dIPβ and the heterologous rescue fails. Created with BioRender.com (accessed on 23 February 2022).
Ijms 23 02613 g001
Figure 2. Scenarios for issues of heterologous rescue results. (A) D. melanogaster wild-type protein dPOIWT encoded by a dGOI has a local interactome, represented by the interacting proteins dIP1, dIP2 and dIP3. When these four proteins interact correctly, the resulting functional complex supports a normal phenotype (top). The functional domain of dPOIWT is D1, which is also present in different hPOIWT equivalents (bottom); some of these equivalents have additional functional domains D2 and D3. (B) hPOIWT allows heterologous rescue (top) or has a steric conformation leading to a loose interaction with dIP2. This condition either prevents formation of the functional complex (no heterologous rescue) or leads to an unstable complex, which determines a partial heterologous rescue result. (C) The normal pathway in humans involves supplemental interactions among hPOIWT, hIP1 and hIP2 (top); D1 of hPOIWT may rescue the phenotype (bottom), but if D2 is defect in patients, chemical repair of D1 in fruit flies drives to a positive heterologous result which cannot be extrapolated to patients. (D) A distinct way to ensemble the functional complex in humans, when the interactome was not evolutionary conserved (upper). hPOIWT has supplemental domains D3 and D4; in this case, D4 domain sterically prevents the proper interaction of D1 with dIP1, therefore heterologous rescue intrinsically fails (bottom). Created with BioRender.com (accessed on 23 February 2022).
Figure 2. Scenarios for issues of heterologous rescue results. (A) D. melanogaster wild-type protein dPOIWT encoded by a dGOI has a local interactome, represented by the interacting proteins dIP1, dIP2 and dIP3. When these four proteins interact correctly, the resulting functional complex supports a normal phenotype (top). The functional domain of dPOIWT is D1, which is also present in different hPOIWT equivalents (bottom); some of these equivalents have additional functional domains D2 and D3. (B) hPOIWT allows heterologous rescue (top) or has a steric conformation leading to a loose interaction with dIP2. This condition either prevents formation of the functional complex (no heterologous rescue) or leads to an unstable complex, which determines a partial heterologous rescue result. (C) The normal pathway in humans involves supplemental interactions among hPOIWT, hIP1 and hIP2 (top); D1 of hPOIWT may rescue the phenotype (bottom), but if D2 is defect in patients, chemical repair of D1 in fruit flies drives to a positive heterologous result which cannot be extrapolated to patients. (D) A distinct way to ensemble the functional complex in humans, when the interactome was not evolutionary conserved (upper). hPOIWT has supplemental domains D3 and D4; in this case, D4 domain sterically prevents the proper interaction of D1 with dIP1, therefore heterologous rescue intrinsically fails (bottom). Created with BioRender.com (accessed on 23 February 2022).
Ijms 23 02613 g002
Table 1. Successful examples of heterologous rescue experiments related to neurodegeneration. Within the vertebrate gene column, (h) indicates a human gene while (m) stands for a mouse gene. Unless otherwise indicated, WT alleles are implicitly considered. HR is the acronym for heterologous rescue and indicates that the references designate HR studies.
Table 1. Successful examples of heterologous rescue experiments related to neurodegeneration. Within the vertebrate gene column, (h) indicates a human gene while (m) stands for a mouse gene. Unless otherwise indicated, WT alleles are implicitly considered. HR is the acronym for heterologous rescue and indicates that the references designate HR studies.
Clinical ImpactVertebrate GeneFly GeneMutant Phenotype (Fly)Heterologous RescueHR
References
motor neuron diseases(h)VAPBVap33loss of Vapp33 determines larval lethality, with few adult escapersexpression of (h)VAPB alleviates the lethal phenotype determined by loss of Vap33[68]
Huntington’s disease(h)UCP2UCP5expression of mutant Huntingtin protein in glia determines altered locomotor performances and uncommon vulnerability to mechanical stressco-expression of (h)UCP2[69]
(h)HTThtthtt-null flies have severe thorax muscle loss and accelerated deterioration in mobility and lifespan(h)HTT rescues the htt loss associated phenotypes[70]
PD(h)MIC60MIC60MIC60mut-null allele determines pupal lethality in homozygous individuals; MIC60mut/+ flies are normalexpression of (h)MIC60 in MIC60mut/+ flies provides a normal phenotype, while expression of mutant (h)MIC60A4V, T11A or C17F leads to severe adult lethality and reduced larval crawling[71]
(h)PRKNparkpark-null flies exhibit reduced lifespan, locomotor and fly defects, infertility, lower cell size and number, progressive degeneration of certain DA neuronsco-expression of (h)PRKN rescues the neurotoxicity; muscle-specific expression of (h)PRKN rescues the flight ability[72,73]
(h)LRRK2LrrkLrrk-null mutants elicit autophagy defects and DA degenerationoverexpression of (h)LRRK2 rescues the mutant phenotype[74,75]
(h)VPS35Vps35downregulation of Vps35 in brain determines supernumerous neuroblast phenotypeexpression of (h)VPS35 fully rescues the brain tumor phenotype exhibited by Vps35 mutants[76]
PD; frontotemporal dementia(h)MAPTtauloss of tau determines lethality; deletion of tau in neurons determines neurodegenerationexpression of (h)MAPT partially rescues the neurodegenerative phenotype[77]
ALS(h)PFN1chicRNAi-mediated downregulation of chic in motor neurons determines pupal lethalitythe chic mutant phenotype is rescued by expressing (h)PFN1 in motor neurons[78]
ALS and other neurodegenerative diseases(h)VCPTER94TER94 mutations determine tubular lysosome dysfunctionexpression of (h)VCP rescues the phenotype determined by mutant TER94[79]
late-onset AD(h)TM2D3amxstrong neurogenic phenotype when amx is maternally mutated(h)TM2D3 is able to partially rescue the neurogenic phenotype and embryonal lethality[80]
Parkinsonism with spasticity, X-linked; intellectual developmental disorder, X-linked, syndromic, Hedera type(h)ATP6AP2ATP6AP2ATP6AP2 depletion is lethal; RNAi knockdown of ATP6AP2 in wing pouch leads to abnormal wing development and growth defectsexpression of (h)ATP6AP2 in ATP6AP2 RNA1 background rescues the specific mutant phenotype[81]
pediatric-onset neurodegenerative disorder(h)ADPRHL2PargParg LOF determines decreased survival in response to oxidative challengelethality is rescued by expressing (h)ADPRHL2[82]
neurodegeneration(h)TARDBPTBPHTBPH-null mutants experience loss of the ventral nerve cord neurons (bursicon neurons)expression of (h)TARDBP rescues the bursicon neurons[83]
neurodegeneration; cancer; metabolic disorder(h)TOMM70Tom70Tom70-null mutation conducted to pupal lethalitythe lethality is rescued by the expression of (h)TOMM70[84]
neurodegeneration; Boucher–Neuhäuser, Gordon Holmes, Laurence–Moon and Oliver McFarlane syndromes(h)PNPLA6swsthe sws1-null mutation causes locomotion deficits and neurodegeneration(h)PNPLA6 rescues the mutant sws phenotype[85]
the sws1 mutants showed characteristic vacuoles in central brain and optic lobes(h)PNPLA6 partially rescues the vacuolization of mutant sws[86]
pantothenate kinase-associated neurodegeneration(h)PanK2fbla hypomorphic mutation in fumble results in flies that have brain lesions, defective neurological functions and severe motor impairmentthe paralysis and impaired climbing activity are rescued by expressing (h)PanK2[87]
in mice, neonatal lethality, slow progressive neurodegeneration, enhanced limb-clasping reflexes, impaired motor activity, cognitive deficits and hypomyelination [88](h)NRD1Nrd1LOF allele causes neurodegenerationexpression of (h)NRD1 rescues the pupal lethality and electroretinogram defects[89]
chorea-acanthocytosis, neurodegeneration, progressive loss of cognitive and locomotor functions(h)VPS13AVps13mutant flies have age-linked neurodegeneration and reduced lifespanoverexpression of (h)VPS13A in mutant flies rescues the characteristic phenotype[90,91]
Alkuraya-Kucinskas and Oliver Mcfarlane syndromes(h)DENND4ACrabflies lacking Crab activity experience age-dependent decline in photoreceptor function and structural integrityexpression of (h)DENND4A rescues the eye defects exhibited by the mutant flies[92]
neurodegenerative encephalopathy(h)TBCDTBCDprojection neurons expressing TBCD1 mutant allele have affected axonal branchesoverexpression of (h)TBCD extensively suppresses the axonal mutant phenotype[93]
neuronal K+–Cl cotransporter; epilepsy(h)SLC12A5kcckccDH1 hypomorphic allele acts as a seizure-enhancer mutation and exacerbates the bang-sensitive paralytic behavior(h)KCC2 rescues the mutant phenotype induced by kccDH1[94]
progressive myoclonus epilepsy(h)GOSR2membrinhomozygosity for membrin-null allele causes larval lethalityexpressing (h)GOSR2 fully rescues the larval lethality, but the adults, although normal looking, display severe motor impairments[95]
early infantile epileptic encephalopathy (EIEE)(h)ACTL6B (BAF53B)Bap55mutations in Bap55 affect the synaptic connections in olfactory neurons(h)ACTL6B rescues the mutant phenotype of Bap55-null individuals[96]
photosensitive epilepsy (PSE)(h)SGMS1-cpes-null mutants show compromised ceramide phosphoethanolamine synthase and fail to complete neuronal cell body encapsulation in the neuronal cortexexpression of (h)SGMS1 rescues the PSE and cortex glial aberrations[97]
ASD(h)TaoK2Taoloss of Tao determines overgrowth of dendritic branching and behavioral defects(h)TaoK2 restores the aberrant dendritic branches to control levels[98,99]
(h)DAT (SLC6A3)DATDAT KO flies are hyperactiveDAT KO flies expressing (h)DAT have reduced locomotion[100,101]
(h)SCAMP1, (h)SCAMP5ScampScamp-null flies exhibit shortened lifespan, compromised climbing, heat-induced seizures and compromised learning and long-term memoryboth (h)SCAMP1 and (h)SCAMP5 rescue the climbing mutant phenotype; (h)SCAMP1 significantly improves the learning index of Scamp-null flies[102]
autism; multiple myeloma(m)Nbeargrg-null mutants exhibit aberrant associative odor learning, modification of gross brain morphology and of synaptic architecturethe transgene (m)Nbea is able to rescue only aversive odor learning and synaptic architecture[103]
affected development of distinct cell types in the central nervous system and in sensory systems(m)Math1atomutant ato embryos lack precursor cell selection and chordotonal organ specificationexpression of (m)Math1 under the control of the ato embryonic enhancer[104]
in mouse, (m)Math1-null animals do not succeed to initiate respiration and die soon after birthreplacing (m)Math1 coding region with ato allowed the animals to survive to adulthood
CMT type 2A, axon degeneration [105](h)MFN1, (h)MFN2Marfmutant flies have affected mitochondria and, as a consequence, their nerves cannot send out signals to muscles; in addition, Marf is lost in the ring gland affecting the production of a hormone required for larva transition to adult, the mutants dying in their larval stageexpression of both (h)MFN1 and (h)MFN2 is necessary for hormone production and the rescue of all phenotypes[106]
CMT neuropathy(h)GDAP1Gdap1knockdown mutants experience retina and muscle degenerationthe mutant phenotype is rescued by (h)GDAP1[107]
dominant-intermediate CMT neuropathy(h)YARSTyrRSRNAi-silenced TyrRS determines specific bristle phenotypesexpressing (h)YARS rescues the abnormal bristle phenotype[108]
CMT neuropathy type 2D(h)GARSGlyRSGlyRS-null flies lack dendritic and axonal terminal arborization(h)GARS rescues the arborization defects in GlyRS-null flies[109]
autosomal recessive cerebellar ataxia(h)UBA5Uba5Uba5-null mutants have reduced lifespan and locomotor activity as well as neuromuscular junction (NMJ) defects(h)UBA5 expression significantly rescues the NMJ mutant phenotype[110]
FA(h)FXNfhfh mutants have altered mitochondrial functions and exhibit age-dependent neurodegenerationexpression of (h)FXN rescues the neurodegeneration[111]
ataxia determined by defects of autophagy(h)ATG5Atg5flies lacking Atg5 activity are unable to walk and fly properly(h)ATG5 restores the mutant flies’ normal movements; (h)ATG5E122D slightly improves the defective mobility[112]
X-linked Snyder–Robinson syndrome(h)SMSSmsSms mutants have critically lowered transcript levels that reduce viability(h)SMS rescues the viability of mutant flies[113]
Delpire–Mcneill syndrome(h)SCL12A2 (NKCC1)Ncc69Ncc69 mutants reach adulthood but their abdominal nerves are swelled and form bulgesthis neuropathy is rescued by (h)SCL12A2[114,115]
microcephaly; Zika virus target(h)ANKLE2Ankle2mutations in Ankle2 can lead to loss of peripheral nervous system organs in adults and severely reduced brain size in hemizygous third instar larvaeexpression of (h)ANKLE2 rescues the mutant phenotype[116,117]
neural network formation; tumor progression(m)BsgBsgmutations in Bsg alter the cell architecture and can lead to high embryo or larval lethalityBsg LOF in adults’ eyes determines mislocalization of photoreceptor nuclei, a phenotype rescued by expressing (m)Bsg[118]
global developmental disorders, intellectual disability(h)CAPZA2cpacpa-null allele determines first instar lethality(h)CAPZA2 rescues the lethal phenotype of cpa-null individuals[119]
autosomal recessive, nonsyndromic intellectual disability(h)ZC3H14Nab2Nab2-null flies experience developmental and locomotor defects(h)ZC3H14 expressed in neurons rescues the Nab2-null phenotype[120]
Troyer syndrome(h)SPG20spartinloss of spartin is associated with motor dysfunctions and brain neurodegenerationsynaptic overgrowth in spartin-null flies is rescued by presynaptic expression of Myc-tagged (h)ZC3H14[121]
intellectual disability(h)OPHN1Grafloss of Graf affects the mushroom body (MB) developmentexpression of (h)OPHN1 significantly ameliorates the MB mutant phenotype[122]
intellectual disability, X-linked(h)CASKCASKaffected expression of CASK negatively impacts middle-term and long-term memoryoverexpression of (h)CASK in neurons of CASK mutants fully rescues the memory[123]
intellectual disability, X-linked(h)ACSL4AcslAcsl mutants exhibit neuromuscular junction overgrowthexpression of (h)ACSL4 rescues the mutant phenotype particular to Acsl mutants[124]
intellectual disability(h)SMARCA5IswiIswi LOF is related to decreased body size and movement in larvae and decreased brain size and locomotor dysfunctions in adults(h)SMARCA5 expression rescues the Iswi specific mutations[125]
nervous system developmental defects(h)EBF3knhomozygous kn-null genotype is embryo lethal(h)EBF3 rescues the lethality[126]
autosomal recessive neurologic disorder(h)TMTC3Tmtc3neuron-specific knockdown of Tmtc3 rises the incidence of mechanically induced seizuresneuron-specific expression of (h)TMTC3[127]
intellectual developmental disorders(h)IQSEC1sizloss of siz affects the growth cones and causes embryonal lethalityoverexpression of (h)IQSEC1 in WT fly background is toxic; lowered expression of (h)IQSEC1 in siz-null mutants partially rescues the embryonal lethality[128]
developmental delay, movement disorders and metabolic decompensation(h)OGDHOgdhLOF allele is associated with early developmental lethalitythe expression of (h)OGDH rescues the mutant phenotype[129]
infantile encephalopathy (lethal)(h)DNM1LDrp1Drp1 mutants have altered mitochondrial trafficking and die as larvaeubiquitous expression of (h)DNM1L rescues the lethality[130]
schizophrenia(h)DTNBP1DysbDysb mutants have compromised memory, elevated climbing activity, abnormal male-male courtship behavior, hypoglutamatergic and hyperdopaminergic activitiespan-neuronal or glial expression of (h)DTNBP1 rescues various Dysb mutant phenotypes[131]
Pitt–Hopkins syndrome(h)TCF4-A, (h)TCF4-Bdada-null allele severely impacts the embryonic nervous system developmentboth (h)TCF4-A and (h)TCF4-B rescue the mutant embryo phenotype[132]
neurofibromatosis, type 2(h)NF2MerMer-null mutations determine lethalityisoform 1 of (h)NF2 is able to rescue the lethality of Mer-null mutants[133]
Table 2. Successful heterologous rescue experiments related to heart disease. Within the vertebrate gene column, (h) indicates a human gene, while (m) stands for a mouse gene. Unless otherwise indicated, WT alleles are implicitly considered. HR is the acronym for heterologous rescue and indicates that the references designate HR studies.
Table 2. Successful heterologous rescue experiments related to heart disease. Within the vertebrate gene column, (h) indicates a human gene, while (m) stands for a mouse gene. Unless otherwise indicated, WT alleles are implicitly considered. HR is the acronym for heterologous rescue and indicates that the references designate HR studies.
Clinical ImpactVertebrate GeneFly GeneMutant Phenotype (Fly)Heterologous RescueHR
References
cardiac dysfunction (postulated), TRiC/CCT complex(h)CCT4CCT4RNAi-silenced CCT4 determines pupal lethality and growth defectsoverexpression of (h)CCT4 rescues the mutant phenotype[187]
lipotoxic cardiomyopathy, ceramide/sphingolipid-related(h)DEGS1ifcknockout of ifc results in
larval lethality
(h)DEGS1 rescues the
lethal phenotype of ifc null
individuals
[188]
congenital heart defect (postulated), KMT2-related(h)KMT2A
(MLL)
trxLOF mutations determines larval to pupal lethality associated with aberrant cuticular patternsexpression of (h)MLL partially rescues the cuticular phenotype[189]
dilated cardiomyopathy 3B(m)DmdDysloss of Dys function leads to reduced lifespan, significantly increased heart rate, age-dependent myofibrillar disorganization, cardiac chamber enlargement and impaired systolic functionthe mutant phenotype was partially reversed by expression of a truncated (m)Dmd
which restores the cardiac diameters and function
[173,178]
Noonan syndrome(h)PTPN11
(SHP-2)
cswcsw mutations determine zygotic lethalityexpression of (h)SHP-2 rescues the zygotic lethality[190]
muscle and aortic defects, ARIH1-related(h)ARIH1ari-1ari-1-null allele is associated with affected larval muscle, lethality or reduced lifespan in adults(h)ARIH1 rescues ari-1-related lethality.[191]
Table 3. Positive heterologous rescue experiments related to cancer. Within the vertebrate gene column, (h) indicates a human gene. Unless otherwise indicated, WT alleles are implicitly considered. HR is the acronym for heterologous rescue and indicates that the references designate HR studies.
Table 3. Positive heterologous rescue experiments related to cancer. Within the vertebrate gene column, (h) indicates a human gene. Unless otherwise indicated, WT alleles are implicitly considered. HR is the acronym for heterologous rescue and indicates that the references designate HR studies.
Clinical ImpactVertebrate GeneFly GeneMutant Phenotype (Fly)Heterologous RescueHR
References
epithelial cancer(h)LLGL1 and (h)LLGL2l(2)gll(2)gl/l(2)gl genotype determines lethality(h)LLGL1 partially rescues the homozygous l(2)gl lethal phenotype; imaginal tissues do not show any neoplastic features, with Dlg and Scrib exhibiting the correct localization; animals undergo a complete metamorphosis and hatch as viable adults[197]
(h)HUGL-1lglmutations in lgl determine structural defects in larvae(h)HUGL-1 expression in the homozygous lgl mutants leads to a partial development of rudimental eyes and larval structures comparable to wild type
(h)LATS1 and (h)LATS2Wtsdevelopmental defects, lethality in flies(h)LATS1 rescues all developmental defects including embryonic lethality in flies[198]
(h)Scribscribpolarity
and neoplastic overgrowth defects
(h)Scrib rescues the polarity
and neoplastic overgrowth defects of scrib mutants
[199]
(h)JRK/JH8ebd1muscle defects in ebd1 and ebd1/ebd2 double mutants(h)JRK/JH8 has rescued the flight muscle defects in ebd1 as well as in ebd1/ebd2 double mutants[200]
(h)CC2D1A and (h)CC2D1BLgdtissue hyperplasia in Lgd mutant phenotype(h)CC2D1A and (h)CC2D1B rescue the Lgd mutant phenotype[201]
various cancers(h)TP53p53p53-null embryos have high sensitivity to genotoxic stressors such as irradiation(h)TP53 partially rescues the embryo liability to irradiation[202]
acute myeloid leukemia(h)MLF1 and (h)MLF2MlfMlf LOF phenotypes include the decrease in embryonic crystal cell numbers and adult bristle and wing phenotypesexpression of (h)MLF1 and (h)MLF2 rescues several Mlf LOF phenotypes[203]
(h)CUX1ctct deficient flies exhibit abnormal wing phenotypes(h)CUX1 rescues the ct mutant phenotypes[204]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ecovoiu, A.A.; Ratiu, A.C.; Micheu, M.M.; Chifiriuc, M.C. Inter-Species Rescue of Mutant Phenotype—The Standard for Genetic Analysis of Human Genetic Disorders in Drosophila melanogaster Model. Int. J. Mol. Sci. 2022, 23, 2613. https://doi.org/10.3390/ijms23052613

AMA Style

Ecovoiu AA, Ratiu AC, Micheu MM, Chifiriuc MC. Inter-Species Rescue of Mutant Phenotype—The Standard for Genetic Analysis of Human Genetic Disorders in Drosophila melanogaster Model. International Journal of Molecular Sciences. 2022; 23(5):2613. https://doi.org/10.3390/ijms23052613

Chicago/Turabian Style

Ecovoiu, Alexandru Al., Attila Cristian Ratiu, Miruna Mihaela Micheu, and Mariana Carmen Chifiriuc. 2022. "Inter-Species Rescue of Mutant Phenotype—The Standard for Genetic Analysis of Human Genetic Disorders in Drosophila melanogaster Model" International Journal of Molecular Sciences 23, no. 5: 2613. https://doi.org/10.3390/ijms23052613

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