Next Article in Journal
nWASP Inhibition Increases Wound Healing via TrKb/PLCγ Signalling
Next Article in Special Issue
Caenorhabditis elegans as a Model System to Study Human Neurodegenerative Disorders
Previous Article in Journal
Opposing Roles of FACT for Euchromatin and Heterochromatin in Yeast
Previous Article in Special Issue
Exogenous Players in Mitochondria-Related CNS Disorders: Viral Pathogens and Unbalanced Microbiota in the Gut-Brain Axis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondrial Neurodegeneration: Lessons from Drosophila melanogaster Models

by
Michele Brischigliaro
1,2,
Erika Fernandez-Vizarra
1,2,* and
Carlo Viscomi
1,2,3,*
1
Department of Biomedical Sciences, University of Padova, 35131 Padova, Italy
2
Veneto Institute of Molecular Medicine, 35129 Padova, Italy
3
Centre for the Study of Neurodegeneration (CESNE), University of Padova, 35131 Padova, Italy
*
Authors to whom correspondence should be addressed.
Biomolecules 2023, 13(2), 378; https://doi.org/10.3390/biom13020378
Submission received: 25 January 2023 / Revised: 9 February 2023 / Accepted: 13 February 2023 / Published: 16 February 2023
(This article belongs to the Special Issue Mitochondria and Central Nervous System Disorders II)

Abstract

:
The fruit fly—i.e., Drosophila melanogaster—has proven to be a very useful model for the understanding of basic physiological processes, such as development or ageing. The availability of straightforward genetic tools that can be used to produce engineered individuals makes this model extremely interesting for the understanding of the mechanisms underlying genetic diseases in physiological models. Mitochondrial diseases are a group of yet-incurable genetic disorders characterized by the malfunction of the oxidative phosphorylation system (OXPHOS), which is the highly conserved energy transformation system present in mitochondria. The generation of D. melanogaster models of mitochondrial disease started relatively recently but has already provided relevant information about the molecular mechanisms and pathological consequences of mitochondrial dysfunction. Here, we provide an overview of such models and highlight the relevance of D. melanogaster as a model to study mitochondrial disorders.

1. Drosophila melanogaster as a Model Organism to Study Disease

The fruit fly (Drosophila melanogaster) has been widely used as a model for research in different fields of biology. The main advantages are its short life cycle (Figure 1); small size; rapid reproductive rate, which is useful for empowering statistical analyses; and the possibility of easy and cheap maintenance of many strains in a limited space.
These features make Drosophila melanogaster an attractive organism for both basic and applied genetics studies. As approximately 75% of human disease-related genes have a functional homolog in the fruit fly genome [1,2], it also constitutes a good model for the study of human disorders. A particular advantage of using D. melanogaster as a model system is access to powerful genetic tools (Figure 2).
For example, transposable P-elements and chemical/physical mutagenesis have been used to induce a large number of mutations and deletions [3,4]. P-elements are transposons that have been engineered to induce genetic modifications through insertional mutagenesis. In addition, the heterologous UAS/GAL4 dual system from S. cerevisiae has been transferred to D. melanogaster in order to finely control the spatiotemporal expression and knockdown of any gene [5,6,7]. This makes it possible to mimic hypomorphic mutations and to overcome frequent experimental limitations linked to lethal or severe phenotypes associated with the complete genetic knockout of essential genes. In fact, the GAL4 transcriptional activator can be expressed under the control of tissue-specific or stage-specific promoters, inducing the expression of elements such as transgenes or inverted-repeats (i.e., single-stranded sequences of nucleotides followed downstream by their reverse complement) for gene knockdown, located downstream of the UAS. Moreover, tissue-specific manipulations are valuable tools that can be used to determine the contribution of each tissue to a disease phenotype. Finally, precise genome-editing techniques developed in the last decade, such as TALENs [8], the CRISPR/Cas9 system [9,10,11], and, more recently, base editing and prime editing [12,13], have been successfully applied to modify the Drosophila genome [14].
Thus, the use of D. melanogaster models provides a rather large array of genetic tools for the investigation of the molecular pathogenesis of human diseases.

2. Mitochondrial Diseases

Mitochondria are double-membrane organelles responsible for the production of most of the ATP in cells via the process of oxidative phosphorylation (OXPHOS) (Figure 3).
The OXPHOS system is composed of four respiratory complexes (complexes I–IV) and two electron carriers (coenzyme Q and cytochrome c), through which an electron funneling cascade coupled with proton pumping allows the generation of an electrochemical gradient across the inner mitochondrial membrane. The electrochemical gradient generates a proton-motif force (pmf) that is exploited by a fifth complex (complex V, ATP synthase) to synthesize ATP from ADP and inorganic phosphate (Pi). ATP stores energy within its phosphodiester bonds, which is then released through the hydrolysis of the bond between the b and g phosphates, driving practically all endergonic biological processes. Although ATP synthesis is often considered the main function of mitochondria, these organelles are key components of many other cellular and metabolic pathways, such as the tricarboxylic acid (TCA or Krebs) cycle, fatty acid oxidation, steroid and pyrimidine synthesis, and the urea cycle. In addition, mitochondria play pivotal roles in several processes, including apoptosis, mitophagy, and intracellular calcium homeostasis.
Mitochondrial diseases are the most frequent inborn errors affecting metabolism, with an estimated prevalence of between 5 and 15 cases per 100,000 individuals [15]. Although extremely heterogeneous from the clinical, biochemical, and genetic points of view, these disorders are all characterized by a dysfunctional OXPHOS system [16], which leads, in most cases, to neurological impairment [17]. More than 340 different genes have been described as being causative of mitochondrial disorders [16]. Mitochondria are peculiar organelles in eukaryotic cells because they contain their own genome, the mitochondrial DNA (mtDNA), which encodes for core components of the OXPHOS complexes (Figure 3). The other subunits of the OXPHOS machinery, as well as all the proteins necessary for its assembly and for the expression of the mitochondrial subunits, originate from nuclear DNA (nDNA), are synthesized in the cytosol, and are actively imported inside the organelle. Thus, mitochondrial diseases can arise from mutations in genes localized in either genome and the inheritance pattern can be either autosomal or X-linked for mutations in nDNA or maternal for mutations in mtDNA. The causes of mitochondrial disease can be classified according to the function of the product of the mutated gene [16] and include not only defects in mtDNA maintenance, mitochondrial gene expression, and synthesis of enzymatic cofactors but also in mitochondrial dynamics and quality control. However, a prominent group of genes associated with mitochondrial disease are those encoding the structural components of the OXPHOS complexes and of specific assembly factors, which are not part of the mature structures but are essential for their proper maturation [18].

3. Models of Mitochondrial Disease in Drosophila melanogaster

3.1. Fly Models of Complex I Defects

Complex I is the largest and most intricate of the respiratory chain enzymes. In humans, it is composed of 44 subunits, 14 of which are “core subunits”—i.e., conserved through evolution from bacteria to humans—while the rest are “supernumerary subunits”, not directly involved in catalysis but important for the stability and/or biogenesis of the enzyme [19,20]. Complex I deficiency is the most common OXPHOS defect and the majority of patients present with neurological impairment, often in the form of Leigh syndrome, with or without the involvement of other organs [21]. Mutations in all mtDNA-encoded subunits, as well as in 24 nuclear-encoded subunits, have been linked to human disease [18].
D. melanogaster has been used as a model to study complex I biogenesis [22,23,24,25], and, importantly, the structure of D. melanogaster complex I has recently been determined [26,27]. D. melanogaster complex I is a 43 subunit complex with high structural homology to its mammalian counterpart and basically the same subunit composition, except for NDUFC1, which is mammalian-specific, and NDUFA2, a highly conserved subunit in terms of sequence but which appears to be loosely associated with the fly complex I [26,27].
Several D. melanogaster models for complex I deficiency have been produced and characterized, mostly using RNAi (Table 1).
These include models for genes encoding both core [28,29,30,31] and supernumerary subunits [31,32,33,34]. Similar to findings reported in humans, flies subjected to knockdown (KD) for ND75 (NDUFS1 homolog) exhibited severe neurological impairment with reduced neuromotor function and longevity [28]. An interesting observation obtained by using cell type-specific KD is that neuronal degeneration is linked to complex I defects in glia rather than primary dysfunction in neurons [28]. Similar findings were reported using a KD model for ND23, the NDUFS8 ortholog, which also unraveled the predominant involvement of glia in the neurodegenerative process [30], despite the lack of behavioral alterations. The specific role of glia in neurological manifestations of mitochondrial disorders has not been investigated in detail in other animal models and patients, but several lines of evidence point to the importance of correct mitochondrial function in glia for neuronal physiology and survival [35].
Fly models of complex I deficiency also include a triple amino acid deletion (p.Met186_Ser188del) in the mtDNA-encoded subunit gene mt:ND2 in the homoplasmic state (mt:ND2del1) [29] obtained by manipulation of the mtDNA with mitochondria-targeted restriction enzymes [36]. Considering the difficulties of manipulating mammalian mtDNA, this and the other mutants for the mtDNA-encoded COX subunits (described below) are extremely relevant models of mtDNA-linked disease [37]. Notably, the mt:ND2del1 variant is not lethal and manifests as a hypomorphic mutation, causing mild neuromotor dysfunction and minor neurodegeneration [29]. However, this model has given important clues about complex I function, as the proton pumping activity is impaired without majorly impacting electron transfer [29].
In addition to defects in genes encoding structural subunits of complex I, other D. melanogaster models with defects in accessory proteins involved in complex I biogenesis (assembly factors) have been characterized, such as CIA30/NDUFAF1 and Sicily/NDUFAF6 [33,38].
These models were generated using different approaches, such as transposon mobilization, chemical mutagenesis, and the UAS/GAL4 system. Even if different genes were targeted, all of them exhibited similar phenotypic features, mainly because they commonly resulted in complex I enzymatic defects. For example, loss of function mutations and strong ubiquitous knockdowns were mostly characterized by severe phenotypes, such as developmental arrest at the larval/pupal stages. On the other hand, hypomorphic mutations and tissue-specific or mild knockdowns usually led to milder phenotypes, often resembling the clinical features observed in patients; i.e., shorter lifespan, decreased neuromotor function, neurodegeneration, seizures, myopathy, increased susceptibility to exogenous stressors, and cardiac dysfunction.

3.2. Fly Models of Complex II Defects

In humans, pathological variants in the four CII structural subunits (SDHA-D) and in two assembly factors (SDHAF1 and SDHAF2) have been associated with either familial tumors, such as paraganglioma and pheochromocytoma, or classical mitochondrial disease [18,39,40]. To date, fly models of three of the four structural subunits of succinate dehydrogenase (SDHA, SDHB, and SDHC) have been produced [41,42,43,44,45]. In addition, mutants in two assembly factors named the Sdhaf3 and Sirup/Sdhaf4 homologs, respectively [45,46], have been characterized (Table 1). Complex II deficiency in flies causes typical mitochondrial dysfunction-associated neurological phenotypes, such as diminished climbing ability, abnormal wing posture, and neurodegeneration, as well as reduced lifespan. Notably, a feature that is frequently observed in complex II deficiencies in flies is hypersensitivity to O2 and increased susceptibility to oxidative stress, with subsequent oxidative damage to proteins [41,43,44,45,46]. An important difference between humans and flies is that, while SDH variants have often been linked to different forms of malignant paragangliomas in humans [47], no evidence has been reported in Drosophila models of CII defects.

3.3. Fly Models of Complex III Defects

Among the OXPHOS defects in human patients, complex III deficiency is the rarest [48]. Using different genetic approaches (i.e., gene KD and KO), three models of complex III deficiency targeting the fly homologues of TTC19, BCS1L [49,50,51], and UQCR10 [52] have been generated and characterized (Table 1). BCS1L and TTC19 pathological variants constitute the most frequently found genetic defects in mitochondrial disease associated with isolated complex III deficiency [48]. Similarly to humans, Ttc19 defects in flies cause a chronic, non-lethal form of neurological CIII deficiency [49,50]. In contrast, Bcs1 knockdown has rather severe effects on D. melanogaster development, as individuals arrest at the larval stage without growing, most probably due to severe CIII deficiency [51]. The fact the partial loss (KD) of the gene causes such a strong phenotype in D. melanogaster is compatible with the fact that, so far, only missense and no loss-of-function mutations have been reported in BCS1L-linked human disease [48].
Notably, the brain-specific silencing of Bcs1 in D. melanogaster allows the larvae to grow and pupate. However, most of the flies die at the pupal stage and, if some individuals survive to adulthood, they suffer from severe paralysis and die in a few days. In contrast, the specific silencing in skeletal muscle leads to complete lethality at the pupal stage [51]. It is important to note that growth retardation, aminoaciduria, cholestasis, iron overload, lactic acidosis, and early death (GRACILE) syndrome is a very severe autosomal recessive human condition linked to one specific BCS1L mutation (p.Ser78Gly) [53]. In contrast to other BCS1L pathological variants, liver failure seems to be a determinant component for the early-onset lethality of GRACILE syndrome [48,54]. However, knockdown of Bcs1 in the fat body, the insect tissue that functionally resembles both the mammalian liver and adipose tissue, has milder effects on D. melanogaster fitness, causing only a slight reduction in lifespan without impacting development [51]. Thus, understanding physiological differences between humans and flies and species-specific features might explain why liver disease is very severe in some specific forms of syndromes caused by BCS1L deficiency.
The third D. melanogaster model of complex III deficiency is linked to a defect in oxen (UQCR10 homolog), a gene that is most likely related to severe cases of in utero onset of ventriculomegaly, apnea, developmental regression, hypotonia, and seizure [55]. Similarly, ox mutants are affected by lethality at the first larval stages [52]. Finally, two neuronal peptides (named sloth1 and sloth2 in D. melanogaster) originating from a bicistronic transcript were linked to complex III biogenesis in flies [56]. Interestingly, sloth1 and sloth2 are homologs of two recently identified mammalian complex III assembly factors named SMIM4 and Brawnin, respectively [57,58]. Complete loss and ubiquitous RNAi of sloth1 and sloth2 cause developmental lethality and neurodegeneration in escaping adults [56].

3.4. Fly Models of Complex IV Defects

Deficiencies in the terminal oxidase of the mitochondrial respiratory chain—i.e., cytochrome c oxidase (COX) or complex IV—are a major cause of mitochondrial disease in humans [59]. Isolated COX deficiencies are mostly associated with mutations in a large number of genes encoding COX structural subunits (either mtDNA- or nDNA-encoded) or, most frequently, assembly factors. COX deficiency is also a feature in patients with mutations in genes encoding mitochondrial gene expression factors, such as LRPPRC, a mitochondrial RNA stabilizing factor; TACO1, a specific translational activator of MT-CO1; or even mitochondrial tRNAs and aminoacyl-tRNA-synthetases [59]. Mutations in nucleus-encoded structural subunits were hypothesized to be embryonic-lethal for a long time because none were found until 2008, when mutations in COX6B1 were identified [60]. After that, several other mutations in other nDNA-encoded genes encoding different complex IV structural subunits were described, but the quantity of disease-related genes encoding COX assembly factors outnumbers the former by far. In the case of COX deficiency, the spectrum of clinical presentations is extremely heterogeneous and ranges from encephalopathic syndromes to cardiomyopathies [59,61,62]. The most frequent presentation of COX deficiency is Leigh syndrome, associated with mutations in SURF1, which encodes an assembly factor with a still-unclear function [63]. COX is highly conserved between humans and flies, with all the 14 subunits composing mammalian COX complex being present in flies, including a COX7B ortholog, which was initially thought to be missing in insects [64]. Missense mutations in two of the three mtDNA-encoded COX subunits (i.e., mt:CoI and mt:CoII, the genes encoding the two catalytic subunits) have been described in the homoplasmic state in D. melanogaster [36,65]. Depending on the mutation, flies displayed a wide range of phenotypes, from healthy (silent) mutations (as in the case of the p.Ala302Thr mutation in mt:CoI) to harmful mutations specific to males (leading to male sterility, such as the p.Arg301Leu mutation) and more severe mutations (such as p.Arg301Ser) triggering growth retardation and neurodegeneration. In addition, numerous models of COX deficiency linked to either defects in nuclear DNA-encoded subunits or assembly factors have been generated and characterized (Table 1). In the early 2000s, mutations in the supernumerary subunits COX5A, levy/COX6A, and cype/COX6C [66,67,68] were introduced.
A few years later, RNAi models for genes encoding the subunits COX4, COX5A, COX5B, levy/COX6A, COX6B, cype/COX6C, and COX7A were described [68,69,70], as well as, more recently, COX7B [64]. Most of these models cause very severe pleiotropic phenotypes, often resulting in developmental lethality and systematically causing neurodegeneration when the RNAi is restricted to the central nervous system.
Several homologs of COX assembly factors have been studied in flies (Table 1). These include Ccdc56/Coa3 [71], the single SCO1 and SCO2 homolog Scox [72,73,74], Surf1 [75,76], and the more recently identified genes encoding metazoan-specific assembly factors Coa7 [77] and Coa8 [78].
Compound heterozygous mutations in COA3 have been identified in one human subject presenting neuropathy linked to COX deficiency [79]. Ccdc56/Coa3 is essential in flies because its complete loss hampers development, causing growth arrest at the larval stage [71]. Ubiquitous RNAi with Surf1 in D. melanogaster is also linked to a severe phenotype and developmental arrest [75,76]. The developmental phenotype of Surf1 RNAi flies is also severe when restricted to muscle [75,76]. Even if SURF1 loss-of-function mutations in humans are associated with severe early-onset encephalopathy, neuronal-specific silencing of Surf1 led to a milder phenotype, with normal development and no major signs of neuropathology. However, slightly decreased neuromotor function was still observed in these flies [75].
A similar observation was recently reported for Scox defects, as neuron-specific knockdown seemed to have little effect on D. melanogaster development and behavior whereas glial KD caused severe deterioration of the neuromotor function [80].
Thus, COX deficiency in flies mimics the human phenotype well, ranging from severe manifestation and early death to neurological disorder. Importantly, cholinergic and adrenergic neurons have been demonstrated to be highly sensitive to COX deficiency in flies, whereas dopaminergic neurons are not.
Recently, by using a set of fly models with KD expression of structural COX subunits (cype) and assembly factors (Coa8, Coa3, and Scox), it was demonstrated that COX defects lead to altered cellular homeostasis and compartmentalization of transition metals; in particular, copper [81]. The contributions of these alterations to the pathogenesis of human diseases warrant more investigation.

3.5. Fly Models of Complex V Defects

The majority of patients with complex V deficiency harbor mutations in the mtDNA region encoding the MT-ATP6 subunit, causing two main phenotypes: either Leigh syndrome or neuropathy, ataxia, and retinitis pigmentosa (NARP) syndrome. However, mutations in the other CV mtDNA-encoded subunit, MT-ATP8, have also been described. Only a few cases of nuclear genes have been identified in patients with complex V deficiency, principally associated with encephalopathic syndromes [18]. These can either encode structural subunits, such as ATP5F1A, ATP5F1D, and ATP5F1E, or assembly factors; namely, ATPAF2 and TMEM70, the latter now considered as an assembly factor for both complex V and complex I [82,83].
A point mutation (p.Gly116Glu) in the mitochondrially encoded mt:ATPase6 gene (MT-ATP6 homolog) resulting in complex V deficiency was found in flies [84]. This was a spontaneous mutation that was identified in the homoplasmic state in flies suffering from a maternally inherited neurodegenerative phenotype and shorter lifespan.
More recently, different genetic manipulation approaches have been exploited to study the effects in flies of defects in ATPsynB, encoding subunit b (ATP5PB in humans); ATPsynC (ATP5MC1 homolog); and ATPsynD, encoding subunit d (the ATP5PD homolog) [85,86,87] (Table 1). Different ATPsynC alleles of varying severity were generated via transposon mobilization and chemical mutagenesis [86]. Null alleles were developmentally lethal whereas hypomorphic alleles caused phenotypes ranging from growth retardation and severe lifespan reduction to hypoactivity and neuromotor dysfunction [86].
Ubiquitous knockdown of ATPsynB and ATPsynD genes resulted in growth arrest and developmental lethality before pupation. Notably, sole misexpression of ATPsynB in testes allowed development but severely impaired fertility in males [87]. In this regard, it is important to note that work undertaken using a set of RNAi targeting D. melanogaster MRC subunits demonstrated that ATP synthase defects in the germline impact differentiation through a mechanism that is independent from OXPHOS dysfunction [88]. The mechanistic details, however, warrant future work.

3.6. Coenzyme Q Deficiency Models

Primary coenzyme Q (CoQ) deficiencies constitute a group of mitochondrial diseases caused by mutations in genes encoding some of the enzymes involved in the synthesis pathway of this essential lipid [89]. As with other mitochondrial diseases, coenzyme Q deficiencies are genetically and clinically extremely heterogeneous. However, the involvement of the CNS in this group of disorders is also very prominent. Specifically, encephalopathy and Leigh-like signs are often present in CoQ deficiencies and typically associated with developmental delay, neuromotor dysfunction, and epilepsy [89]. Defective biosynthesis of CoQ in D. melanogaster has been investigated by studying mutations in qless, the PDSS1 ortholog [90] (Table 1). Despite the fact that some forms of CoQ deficiencies also manifest with a renal phenotype, qless mutation in D. melanogaster leads to severe specific defects in the CNS, with increased caspase activation and neuronal death, similarly to most of the human cases reported with mutations in CoQ-related genes [89].
Table 1. Fly models of respiratory chain defect.
Table 1. Fly models of respiratory chain defect.
Fly GeneHuman OrthologFunctionSystemTissue SpecificityPhenotypeRef.
Complex Imt:ND2MT-ND2Core subunitRestriction enzymes targeting mtDNAUbiquitousNeuromotor dysfunction, neurodegeneration[29]
ND-75NDUFS1Core subunitRNAiGliaNeurodegeneration[28]
RNAiUbiquitousNeurodegeneration
RNAiNeuronsReduced lifespan
ND-23NDUFS8Core subunitRNAiGliaNeurodegeneration[30]
RNAiUbiquitousDevelopmental arrest
RNAiNeuronsReduced lifespan, neuromotor dysfunction
ND-20NDUFS7Core subunitRNAiUbiquitousArray of phenotypes depending on RNAi efficiency[31]
ND-51NDUFV1Core subunitRNAiUbiquitousDevelopmental arrest
ND-19NDUFA8Supernumerary subunitRNAiUbiquitousDevelopmental arrest[32]
ND-39NDUFA9Supernumerary subunitRNAiUbiquitousDevelopmental arrest[32]
ND-42NDUFA10Supernumerary subunitRNAiUbiquitousDevelopmental arrest[33]
RNAiEyeRetinal degeneration
SicilyNDUFAF6Assembly factorFLP/FRT systemMosaic eyeRetinal degeneration, neurodegeneration[33]
Transposable elementsUbiquitousDevelopmental arrest
ND-18NDUFS4Supernumerary subunitRNAiUbiquitousArray of phenotypes depending on RNAi efficiency[31]; [34]
CIA30NDUFAF1Assembly factorTransposable elementsUbiquitousDevelopmental arrest[38]
RNAiUbiquitousReduced growth, partial developmental lethality
Complex IISdhASDHASubunitFLP/FRT systemMosaic eyeRetinal degeneration[41]
FLP/FRT systemUbiquitousDevelopmental arrest
SdhBSDHBSubunitTransposable elementsUbiquitousReduced lifespan, sensitivity to hyperoxia, age-related neuromotor dysfunction[43]
SdhCSDHCSubunitOverexpression of dominant negative mutationNeuronalReduced lifespan, oxidative damage[44]
Sirup/Sdhaf4SDHAF4Assembly factorTALENsUbiquitousReduced lifespan, neurodegeneration, sensitivity to oxidative stress[45]
Sdhaf3SDHAF3Assembly factorHomologous recombinationUbiquitousSensitivity to oxidative stress and hyperoxia, age-related neuromotor dysfunction[46]
Ttc19TTC19Assembly factorTransposable elementsUbiquitousNeuromotor dysfunction[49]
CRISPR/Cas9 KOUbiquitousNeuromotor dysfunction[50]
Bcs1BCS1LAssembly factorRNAiUbiquitousDevelopmental arrest, larval neuromotor dysfunction[51]
RNAiNeuronsReduced lifespan, neuromotor dysfunction, paralysis
RNAiMuscleDevelopmental arrest
RNAiFat bodyReduced lifespan
OxUQCR10Supernumerary subunitTransposable elementsUbiquitousDevelopmental arrest[52]
sloth1SMIM4/
UQCC5
Assembly factorRNAiUbiquitousDevelopmental lethality, neurodegeneration[56]
CRISPR/Cas9 KOUbiquitous (somatic)Developmental lethality, neurodegeneration
CRISPR/Cas9 KOUbiquitous (germline)Developmental lethality, neurodegeneration
sloth2Brawnin/
UQCC6
Assembly factorRNAiUbiquitousDevelopmental lethality, neurodegeneration[56]
CRISPR/Cas9 KOUbiquitous (somatic)Developmental lethality, neurodegeneration
CRISPR/Cas9 KOUbiquitous (germline)Developmental lethality, neurodegeneration
Complex IVmt:CoIMT-CO1Core subunitMitochondrially targeted restriction enzymesUbiquitousReduced growth, neurodegeneration[36]
COX7BCOX7BSupernumerary subunitRNAiUbiquitousDevelopmental arrest[64]
cype/
COX6C
COX6CSupernumerary subunitFLP/FRT systemEyeRetinal degeneration[66]
FLP/FRT systemGermlineDevelopmental arrest
RNAiUbiquitousDevelopmental arrest[81]
COX5ACOX5ASupernumerary subunitFLP/FRT systemEyeRetinal degeneration[67]
RNAiUbiquitousDevelopmental arrest[70]
levy/
COX6A
COX6A1Supernumerary subunitChemical mutagenesisUbiquitousTemperature-induced paralysis, bang-induced paralysis, neurodegeneration, reduced lifespan[68]
RNAiUbiquitousDevelopmental lethality[70]
COX4COX4I1Supernumerary subunitRNAiUbiquitousDevelopmental arrest (strong RNAi), reduced lifespan (mild RNAi)[69]
COX5BCOX5BSupernumerary subunitRNAiUbiquitousDevelopmental arrest[69]
RNAiUbiquitousDevelopmental arrest[70]
COX7ACOX7A1Supernumerary subunitRNAiUbiquitousDevelopmental arrest[70]
Ccdc56/Coa3COA3Assembly factorTransposable elementsUbiquitousDevelopmental arrest[71]
ScoxSCO1/
SCO2
Assembly factorTransposable elementsUbiquitousDevelopmental arrest[72]
RNAiUbiquitousDevelopmental arrest[73]
RNAiHeartReduced lifespan, cardiac dysfunction[74]
RNAiGliaNeuromotor dysfunction[80]
Surf1SURF1Assembly factorRNAiUbiquitousDevelopmental arrest[75]; [76]
RNAiNeuronsMild neuromotor defects[75]
RNAiMuscleDevelopmental arrest[76]
Coa7COA7Assembly factorRNAiEyeRetinal degeneration[77]
RNAiNeuronsReduced lifespan, neuromotor dysfunction
Coa8COA8Assembly factorRNAiUbiquitousSensitivity to oxidative stress, neuromotor dysfunction[78]
RNAiNeuronsSensitivity to oxidative stress, neuromotor dysfunction
Complex Vmt:ATPase6MT-ATP6Core subunitIsolation of spontaneous mutationUbiquitousReduced lifespan, progressive neurodegeneration[84]
ATPsynDATP5PDCore subunitRNAiUbiquitousDevelopmental arrest[85]
ATPsynBATP5PBCore subunitRNAiUbiquitousDevelopmental arrest[87]
ATPsynCATP5MC1/ATP5MC2/ATP5MC3Core subunitTransposable elements, chemical mutagenesisUbiquitousRange of phenotypes depending on the severity of the genetic lesion[86]
CoQqlessPDSS1CoQ biosynthesisChemical mutagenesisUbiquitousDevelopmental arrest[90]
FLP/FRT systemNeuronsNeurodegeneration

3.7. Defects in Mitochondrial DNA Replication and Maintenance

So far, numerous genes have been linked to mtDNA replication and maintenance defects in human disease [91]. These include genes encoding factors that are directly dedicated to replication of the mitochondrial genome, such as POLG, POLG2, TWNK, and TFAM, and those indirectly involved in the maintenance of mtDNA, such as enzymes involved in dNTP synthesis (e.g., TK2, DGUOK, SUCLG1, and SUCLA2). Other genes associated with mtDNA instability have unknown functions (e.g., MPV17). It is important to note that mutations in genes encoding proteins involved in mitochondrial dynamics (e.g., OPA1 and MFN2) can also cause mtDNA maintenance disorders, as proper mitochondrial architecture seems to be essential for correct mtDNA replication [92].
Defects in D. melanogaster POLγ, the mtDNA-specific DNA polymerase, were first reported in 1999 [93]. In fact, the gene encoding the catalytic subunit (subunit α) of mtDNA polymerase, initially named tamas (the Sanskrit word for “darkness”)—official symbol PolG1—was identified during a screening of pupal lethal phenotypes. Numerous pathogenic alleles of PolG1 have been described since then, most of them affecting viability at or before the pupal stage [93,94,95] (Table 2).
Importantly, D. melanogaster POLγ has been engineered to generate models making it possible to study the effects of random generation and accumulation of mtDNA mutations in vivo (mtDNA mutator models). Firstly, the exonuclease domain of PolG1 was mutated to impair the proofreading activity of the enzyme, and this mutant was used to complement a PolG1 KO strain [94]. Homozygosity in proofreading defective PolG1 (named the exo allele) causes developmental lethality in D. melanogaster, but heterozygous individuals do not show behavioral defects, despite having increased mutational rates in mtDNA throughout the generations [94]. In addition, a second mutator fly model carrying the very same mutation in the proofreading domain of POLγ was generated using a different approach, which was transgenic expression of exo PolG1 [96]. Notably, in this work, the authors noted some differences between the two mutator fly models. In fact, while the first model was lethal in homozygosity [94], this was not observed in the second model [96]. The causes behind these discrepancies are currently unclear, but they might be explained by differing mutational heterogeneity between the two models. Further, and importantly, mtDNA heteroplasmy levels are likely to have a primary modifying role. In fact, studies using the analogous mutator mouse model have also led to an intense debate regarding the role of mtDNA mutations in disease and aging [97,98,99]. It is worth mentioning that an alternative approach for generating a D. melanogaster mtDNA mutator model was based on mitochondrial targeting of APOBEC1, a vertebrate cytidine deaminase enzyme [100]. This enzyme can specifically introduce point mutations that do not affect the mtDNA copy number, introduce insertions/deletions, or affect development. However, the accumulation of mtDNA mutations did cause early death and mitochondrial dysfunction in the adult stage.
Fly disease models of the mtDNA-helicase gene (mammalian TWNK) have also been generated and studied (Table 2). Firstly, three mtDNA-helicase variants corresponding to human autosomal dominant PEO mutations were expressed in vivo [101]. Two of them (p.Lys388Ala and p.Ala442Pro) caused mtDNA depletion and severe phenotypes, resulting in arrest at different developmental phases before the adult stage. Curiously, the third dominant mutation (p.Trp441Cys) did not show strong effects, as mtDNA depletion levels were minimal and no developmental arrest was observed. In addition, RNAi was used to perturb mtDNA-helicase gene expression [95]. Similar to the effect of the overexpression of dominant negative mutants, KD of the helicase-encoding gene resulted in mtDNA depletion and lethality of around 75% in individuals at the pupal stage.
Recently, variants of bor (belphegor), the homolog of ATAD3A, a gene associated with mitochondrial disease in humans and encoding a component of the nucleoid (i.e., the association of mtDNA and proteins) [102,103,104], have been studied in D. melanogaster [103,105]. In these cases, the phenotypes observed in flies harboring missense pathogenic variant were compatible with mitochondrial disease and included hypotonia, developmental delay, cardiomyopathy, and brain abnormalities [105], whereas complete loss of bor had been previously linked to growth arrest at the larval stage [106].
Mutations in SUCLG1, encoding the alpha subunit of the succinyl-CoA synthetase, cause severe early-onset mtDNA depletion syndromes in humans [107,108,109]. In contrast, loss of function in the fly homolog Scsα1 does not lead to early lethal phenotypes. However, disease phenotypes, such as developmental delay, altered locomotor behavior, and reduced lifespan under starvation, were still observed [110].
Very recently, a neuron-specific RNAi Drosophila model of MPV17 (dMpv17) was reported and showed impaired locomotor activity in larvae and learning ability in adults, altered energy metabolism, and abnormal neuromuscular junctions [111]. This is an interesting observation, as patients, who were characterized by early-onset liver failure due to profound depletion of mtDNA in the liver, developed a progressive neurological phenotype at later stages [112]. In addition, peripheral neuropathy has been reported for some patients [113].

3.8. Defects in Mitochondrial Gene Expression

Mitochondria contain separated gene expression machineries for the synthesis of the mtDNA-encoded polypeptides; i.e., specific mitochondrial transcription and translation factors (nDNA encoded), mtDNA-encoded transfer RNAs (tRNAs), and mitochondrial ribosomes (mitoribosomes) composed of nDNA-encoded proteins and of ribosomal RNAs (rRNAs) encoded in the mtDNA [114]. In the last few years, many factors involved in mitochondrial gene expression—in particular, translation—have been linked to human disorders, including mutations in mitoribosomal proteins [115,116]. Most of the disorders linked to mitochondrial gene expression have neurological manifestations, such as leukoencephalopathy and Leigh syndrome [115,116].
Notably, the first D. melanogaster model of mitochondrial dysfunction was a mitochondrial ribosomal protein mutant. This model was reported in 1987 when a pathological variant of the technical knockout (tko) gene was found in homozygosity in flies suffering from a neurological temporary paralytic phenotype induced by mechanical shock known as “bang sensitivity” [117]. The gene was found to encode the mitochondrial ribosomal protein S12 (mRpS12). Later, the tko fly model was further studied as a model of mitochondrial disease. Indeed, in addition to bang sensitivity, mutant flies were found to suffer from developmental delay, hypersensitivity to doxycycline (an inhibitor of mitochondrial translation), and deafness due to mitochondrial respiratory chain dysfunction [118]. More recently, neuronal-specific RNAi of D. melanogaster genes encoding the mitochondrial ribosomal proteins mRpL15 and mRpL40 showed disruption of synapse development and function [119]. Thus, altered mitochondrial translation also predominantly causes neurological phenotypes in D. melanogaster (Table 2).

3.9. Defects in Mitochondrial Dynamics and Architecture

In recent decades, interest in the influence of mitochondrial architecture and dynamics on health and disease has increased considerably. Mitochondria are not static and isolated organelles but instead form a highly dynamic “mitochondrial network” governed by fission and fusion processes [120]. In fact, proper “mitodynamics” appears to be very relevant for different processes related to mitochondrial function, such as mtDNA replication, metabolism, and recycling of dysfunctional/damaged mitochondria [120]. Moreover, proper mitodynamics is important for organism development [121,122,123,124]. As a consequence, human mitochondrial disorders can also be caused by dysfunctional components controlling fission and fusion, such as OPA1, MFN1-2, MFF, and DRP1 [125].
Along with studies performed in vitro, several animal models, including D. melanogaster models (Table 2), have been generated to study the effects of defective mitochondrial morphology and dynamics in vivo. Several mutations in fly Drp1 lead to developmental lethality at the third larval stage, associated with impaired neurotransmission [126], and the specific loss of Drp1 in spermatocytes leads to altered spermatogenesis due to mitochondrial clustering and impaired motility [127]. Essential factors for maintaining mitochondrial network morphology and cristae shape include OPA1 and MFN2, the dysfunction of which causes an array of phenotypes in D. melanogaster linked to defective mitochondrial architecture, including developmental delay or arrest, cardiomyopathy, and neurological phenotypes resembling autosomal dominant optic atrophy (ADOA) and Charcot–Marie–Tooth type 2A syndrome (CMT2A) [128,129,130,131].
In addition, the inner mitochondrial membrane ultrastructure is intimately related to mitodynamics because it depends on the functions of proteins such as OPA1 or DRP1. However, mitochondrial ultrastructure is also heavily influenced by other factors, such as the dimerization of complex V (ATP synthase) at the cristae rims [132,133] and the mitochondrial contact sites and cristae organization system (MICOS) complex [134].
Mutations in some MICOS components have been reported in humans, including MIC13/QIL1 [135,136,137], causing a severe form of infantile hepato-encephalopathy; MIC26/APOO [138], associated with an X-linked mitochondrial myopathy with cognitive impairment; and MIC60, linked to Parkinsonism [139]. In D. melanogaster, deletion of the main component of the MICOS complex—MIC60/Mitofilin—causes a severe developmental phenotype with growth arrest at the pupal stage [139,140]. However, APOO loss in D. melanogaster is associated with milder phenotypes, partial developmental lethality, and mitochondrial ultrastructure defects with multiple OXPHOS deficiencies [138]. Notably, MIC13/QIL1 depletion in muscle and neurons causes abnormal mitochondrial network, ultrastructure, and function [141].
Finally, a member of the solute carrier family (named SLC25A46) has repeatedly been reported to be associated to different forms of neurological mitochondrial disorder and Leigh syndrome [142,143,144,145]. SLC25A46 encodes a mitochondrial outer membrane protein involved in mitochondrial dynamics that interacts with MFN2, OPA1, and MICOS [143]. A D. melanogaster model for Slc25A46a was recently described [146]. Specifically, Slc25A46a knockdown in fly neurons causes neurological phenotypes both in larvae and adults, with reduced neuromotor function and altered morphology in the neuromuscular junction [146].
Table 2. Fly models of other mitochondrial defects.
Table 2. Fly models of other mitochondrial defects.
Fly GeneHuman OrthologFunctionSystemTissue SpecificityPhenotypeRef.
mtDNA replication and maintenancePolG1/tamPOLGmtDNA replicationChemical mutagenesisUbiquitousDevelopmental arrest, neuromotor dysfunction[93]
Homologous recombinationUbiquitousDevelopmental arrest, reduced growth[94]
RNAiUbiquitousDevelopmental arrest[95]
KI of PolG1 exo (mutator)UbiquitousDevelopmental lethality in homozygosity, increased mtDNA mutation rate in heterozygosity[94]
Transgenic PolG1 exo (mutator)UbiquitousReduced lifespan, dose-dependent increase in mtDNA mutation rate[96]
RNAiUbiquitousPartial developmental lethality[95]
mtDNA-helicaseTWNKmtDNA replicationTransgenic expression of dominant mutationsUbiquitousDevelopmental arrest and mtDNA depletion[101]
borATAD3AComponent of nucleoidsTransgenic expression of dominant mutationUbiquitousDevelopmental arrest[103]
NeuronsDevelopmental arrest
Muscle-specificPartial developmental lethality
Transposable elementsUbiquitousDevelopmental arrest[106]
SCSα1SUCLG1Mitochondrial nucleotide synthesisCRISPR/Cas9UbiquitousDevelopmental delay, altered neuromotor function, and reduced lifespan under starvation[110]
Mitochondrial translationtkoMRPS12Mitoribosome small subunitChemical mutagenesisUbiquitousBang-induced paralysis, developmental delay, sensitivity to doxycycline[117]; [118]
mRpL15MRPL15Mitoribosome large subunitRNAiNeuronsDisruption of synapse development and function[119]
mRpL40MRPL40Mitoribosome large subunitRNAiNeuronsDisruption of synapse development and function[119]
Mitochondrial dynamics/architectureDrp1DRP1Mitochondrial fissionTransposable elementsUbiquitousPartial developmental lethality, altered neuromotor function[126]
FLP/FRT systemSpermatocytesAltered spermatogenesis and sperm motility[127]
Opa1OPA1Mitochondrial fusionFLP/FRT systemEyeRetinal degeneration[128]
Transposable elementsUbiquitousDevelopmental arrest[128]
RNAiHeartCardiomyopathy[131]
MarfMFN1/2Mitochondrial fusionRNAiHeartCardiomyopathy[131]
Mic26-27MIC26-27Cristae architectureTransposable elementsUbiquitousPartial developmental lethality, reduced lifespan, reduced neuromotor function[138]
MitofilinIMMT/ MIC60Cristae architectureTransposable elementsUbiquitousDevelopmental arrest[139]
RNAiMuscleMild neuromotor defects
RNAiNeuronsMild neuromotor defects
Slc25A46aSLC25A46Mitochondrial dynamicsRNAiNeuronsNeuromotor dysfunction[146]

4. Conclusions

The growing number of Drosophila melanogaster models of mitochondrial deficiency underscores their usefulness in the study of the phenotypical, biochemical, and molecular features of human mitochondrial diseases. As we described in this review, in many cases, specific genetic defects leading to OXPHOS deficiency result in observable pathological phenotypes resembling the main clinical features of patients. Notably, while the mouse models of mitochondrial disease often poorly reproduce the neurological signs typical of the human disease, flies usually show neurological phenotypes, and the study of several Drosophila models of mitochondrial dysfunction unraveled the central role of glia in the development of neurological phenotypes. This will open the ground for future investigations to address the pathological role of the glia in mammalian models and mitochondrial disease patients. Therefore, generating and studying these fruit fly strains has provided a key instrument not only for the validation of the pathological significance of the genetic variants found in human patients but also for the understanding of the basic cellular and molecular mechanisms related to mitochondrial diseases. The main advantages of using D. melanogaster models for these investigations are the easy genetic manipulation and short generation times. In fact, genetic knockdown by RNAi, which is easily and routinely applied in flies, provides a system that better resembles the situation of hypomorphic alleles, which is more frequently encountered in human mitochondrial disorders than total KO or loss-of-function mutations. A limitation of D. melanogaster is that, in several cases, mutations associated in humans with post-natal diseases cause developmental arrest in flies, probably due to the high energy requirements and peculiar metabolism during larva-to-pupa and pupa-to-adult transitions.
Furthermore, in practical terms, having established reliable models, D. melanogaster can be used as a valuable and cost-effective—but still complex—animal model that can complement the observations obtained using other models, such as murine models, which are subject to tighter ethical regulations and are much more costly timewise and economically. In addition, mice models frequently enough do not closely recapitulate human diseases. Therefore, due to these reasons, the generation of fly models can facilitate several types of translational studies, such as medium-scale drug screenings, which are necessary in order to find efficient therapies for mitochondrial diseases [147].
In conclusion, ease of handling and the low requirements for equipment and funding to carry out studies make D. melanogaster an attractive system for biomedical research and, more specifically, for investigations into genetic disorders, such as mitochondrial diseases.

Author Contributions

M.B., E.F.-V. and C.V. drafted and finalized the manuscript. E.F.-V. and C.V. acquired funding. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by: Telethon/Cariplo Foundation (grant GJC21014 to E.F.-V.), Department of Biomedical Sciences—UNIPD (FERN_FAR22_01 to E.F.-V.), Telethon Foundation (GGP20013 to C.V.), AFM-Telethon (23706 to C.V.), Associazione Luigi Comini Onlus (MitoFight2, to C.V.), and the Department of Biomedical Sciences—UNIPD (SID2022- VISC_BIRD2222_01 to C.V.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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]
  2. Pandey, U.B.; Nichols, C.D. Human Disease Models in Drosophila Melanogaster and the Role of the Fly in Therapeutic Drug Discovery. Pharmacol. Rev. 2011, 63, 411–436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Bellen, H.J.; Levis, R.W.; Liao, G.; He, Y.; Carlson, J.W.; Tsang, G.; Evans-Holm, M.; Hiesinger, P.R.; Schulze, K.L.; Rubin, G.M.; et al. The BDGP Gene Disruption Project: Single Transposon Insertions Associated with 40% of Drosophila Genes. Genetics 2004, 167, 761–781. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bellen, H.J.; Levis, R.W.; He, Y.; Carlson, J.W.; Evans-Holm, M.; Bae, E.; Kim, J.; Metaxakis, A.; Savakis, C.; Schulze, K.L.; et al. The Drosophila Gene Disruption Project: Progress Using Transposons with Distinctive Site Specificities. Genetics 2011, 188, 731–743. [Google Scholar] [CrossRef] [Green Version]
  5. Brand, A.H.; Perrimon, N. Targeted Gene Expression as a Means of Altering Cell Fates and Generating Dominant Phenotypes. Development 1993, 118, 401–415. [Google Scholar] [CrossRef]
  6. Dietzl, G.; Chen, D.; Schnorrer, F.; Su, K.-C.; Barinova, Y.; Fellner, M.; Gasser, B.; Kinsey, K.; Oppel, S.; Scheiblauer, S.; et al. A Genome-Wide Transgenic RNAi Library for Conditional Gene Inactivation in Drosophila. Nature 2007, 448, 151–156. [Google Scholar] [CrossRef]
  7. Caygill, E.E.; Brand, A.H. The GAL4 System: A Versatile System for the Manipulation and Analysis of Gene Expression. Methods Mol. Biol. 2016, 1478, 33–52. [Google Scholar] [CrossRef]
  8. Liu, J.; Li, C.; Yu, Z.; Huang, P.; Wu, H.; Wei, C.; Zhu, N.; Shen, Y.; Chen, Y.; Zhang, B.; et al. Efficient and Specific Modifications of the Drosophila Genome by Means of an Easy TALEN Strategy. J. Genet. Genom. 2012, 39, 209–215. [Google Scholar] [CrossRef]
  9. Bassett, A.; Liu, J.L. CRISPR/Cas9 Mediated Genome Engineering in Drosophila. Methods 2014, 69, 128–136. [Google Scholar] [CrossRef]
  10. Housden, B.E.; Lin, S.; Perrimon, N. Cas9-Based Genome Editing in Drosophila. Methods Enzymol. 2014, 546, 415–439. [Google Scholar] [CrossRef]
  11. Gratz, S.J.; Rubinstein, C.D.; Harrison, M.M.; Wildonger, J.; O’Connor-Giles, K.M. CRISPR-Cas9 Genome Editing in Drosophila. Curr. Protoc. Mol. Biol. 2015, 2015, 31.2.1–31.2.20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Bosch, J.A.; Birchak, G.; Perrimon, N. Precise Genome Engineering in Drosophila Using Prime Editing. Proc. Natl. Acad. Sci. USA 2020, 118, e2021996118. [Google Scholar] [CrossRef]
  13. Marr, E.; Potter, C.J. Base Editing of Somatic Cells Using CRISPR-Cas9 in Drosophila. CRISPR J. 2021, 4, 836–845. [Google Scholar] [CrossRef] [PubMed]
  14. Zirin, J.; Bosch, J.; Viswanatha, R.; Mohr, S.E.; Perrimon, N. State-of-the-Art CRISPR for in Vivo and Cell-Based Studies in Drosophila. Trends Genet. 2022, 38, 437–453. [Google Scholar] [CrossRef] [PubMed]
  15. Gorman, G.S.; Chinnery, P.F.; DiMauro, S.; Hirano, M.; Koga, Y.; McFarland, R.; Suomalainen, A.; Thorburn, D.R.; Zeviani, M.; Turnbull, D.M. Mitochondrial Diseases. Nat. Rev. Dis. Prim. 2016, 2, 16080. [Google Scholar] [CrossRef]
  16. Gusic, M.; Prokisch, H. Genetic Basis of Mitochondrial Diseases. FEBS Lett. 2021, 595, 1132–1158. [Google Scholar] [CrossRef]
  17. Zeviani, M.; Viscomi, C. Mitochondrial Neurodegeneration. Cells 2022, 11, 637. [Google Scholar] [CrossRef]
  18. Fernandez-Vizarra, E.; Zeviani, M. Mitochondrial Disorders of the OXPHOS System. FEBS Lett. 2021, 595, 1062–1106. [Google Scholar] [CrossRef]
  19. Stroud, D.A.; Surgenor, E.E.; Formosa, L.E.; Reljic, B.; Frazier, A.E.; Dibley, M.G.; Osellame, L.D.; Stait, T.; Beilharz, T.H.; Thorburn, D.R.; et al. Accessory Subunits Are Integral for Assembly and Function of Human Mitochondrial Complex i. Nature 2016, 538, 123–126. [Google Scholar] [CrossRef] [Green Version]
  20. Padavannil, A.; Ayala-Hernandez, M.G.; Castellanos-Silva, E.A.; Letts, J.A. The Mysterious Multitude: Structural Perspective on the Accessory Subunits of Respiratory Complex I. Front. Mol. Biosci. 2022, 8, 1252. [Google Scholar] [CrossRef]
  21. Rodenburg, R.J. Mitochondrial Complex I-Linked Disease. Biochim. Biophys. Acta Bioenerg. 2016, 1857, 938–945. [Google Scholar] [CrossRef] [PubMed]
  22. Garcia, C.J.; Khajeh, J.; Coulanges, E.; Chen, E.I.-j.; Owusu-Ansah, E. Regulation of Mitochondrial Complex I Biogenesis in Drosophila Flight Muscles. Cell Rep. 2017, 20, 264–278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Rhooms, S.K.; Murari, A.; Goparaju, N.S.V.; Vilanueva, M.; Owusu-Ansah, E. Insights from Drosophila on Mitochondrial Complex I. Cell. Mol. Life Sci. 2020, 77, 607–618. [Google Scholar] [CrossRef] [PubMed]
  24. Murari, A.; Rhooms, S.K.; Goparaju, N.S.; Villanueva, M.; Owusu-Ansah, E. An Antibody Toolbox to Track Complex I Assembly Defines AIF’s Mitochondrial Function. J. Cell Biol. 2020, 219, e202001071. [Google Scholar] [CrossRef]
  25. Murari, A.; Rhooms, S.K.; Garcia, C.; Liu, T.; Li, H.; Mishra, B.; Deshong, C.; Owusu-Ansah, E. Dissecting the Concordant and Disparate Roles of NDUFAF3 and NDUFAF4 in Mitochondrial Complex I Biogenesis. iScience 2021, 24, 102869. [Google Scholar] [CrossRef]
  26. Agip, A.-N.A.; Chung, I.; Sanchez-Martinez, A.; Whitworth, A.J.; Hirst, J. Cryo-EM Structures of Mitochondrial Respiratory Complex I from Drosophila Melanogaster. eLife 2023, 12, e84424. [Google Scholar] [CrossRef]
  27. Padavannil, A.; Murari, A.; Rhooms, S.-K.; Owusu-Ansah, E.; Letts, J.A. Resting Mitochondrial Complex I from Drosophila Melanogaster Adopts a Helix-Locked State. bioRxiv 2022. [Google Scholar] [CrossRef]
  28. Hegde, V.R.; Vogel, R.; Feany, M.B. Glia Are Critical for the Neuropathology of Complex I Deficiency in Drosophila. Hum. Mol. Genet. 2014, 23, 4686–4692. [Google Scholar] [CrossRef] [Green Version]
  29. Burman, J.L.; Itsara, L.S.; Kayser, E.B.; Suthammarak, W.; Wang, A.M.; Kaeberlein, M.; Sedensky, M.M.; Morgan, P.G.; Pallanck, L.J. A Drosophila Model of Mitochondrial Disease Caused by a Complex I Mutation That Uncouples Proton Pumping from Electron Transfer. DMM Dis. Model. Mech. 2014, 7, 1165–1174. [Google Scholar] [CrossRef] [Green Version]
  30. Cabirol-Pol, M.J.; Khalil, B.; Rival, T.; Faivre-Sarrailh, C.; Besson, M.T. Glial Lipid Droplets and Neurodegeneration in a Drosophila Model of Complex I Deficiency. Glia 2018, 66, 874–888. [Google Scholar] [CrossRef] [Green Version]
  31. Foriel, S.; Herma Renkema, G.; Lasarzewski, Y.; Berkhout, J.; Rodenburg, R.J.; Smeitink, J.A.M.; Beyrath, J.; Schenck, A. A Drosophila Mitochondrial Complex i Deficiency Phenotype Array. Front. Genet. 2019, 10, 245. [Google Scholar] [CrossRef] [PubMed]
  32. Sanz, A.; Soikkeli, M.; Portero-Otín, M.; Wilson, A.; Kemppainen, E.; McIlroy, G.; Ellilä, S.; Kemppainen, K.K.; Tuomela, T.; Lakanmaa, M.; et al. Expression of the Yeast NADH Dehydrogenase Ndi1 in Drosophila Confers Increased Lifespan Independently of Dietary Restriction. Proc. Natl. Acad. Sci. USA 2010, 107, 9105–9110. [Google Scholar] [CrossRef] [Green Version]
  33. Zhang, K.; Li, Z.; Jaiswal, M.; Bayat, V.; Xiong, B.; Sandoval, H.; Charng, W.L.; David, G.; Haueter, C.; Yamamoto, S.; et al. The C8ORF38 Homologue Sicily Is a Cytosolic Chaperone for a Mitochondrial Complex i Subunit. J. Cell Biol. 2013, 200, 807–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Foriel, S.; Beyrath, J.; Eidhof, I.; Rodenburg, R.J.; Schenck, A.; Smeitink, J.A.M. Feeding Difficulties, a Key Feature of the Drosophila NDUFS4 Mitochondrial Disease Model. DMM Dis. Model. Mech. 2018, 11, dmm032482. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Rose, J.; Brian, C.; Woods, J.; Pappa, A.; Panayiotidis, M.I.; Powers, R.; Franco, R. Mitochondrial Dysfunction in Glial Cells: Implications for Neuronal Homeostasis and Survival. Toxicology 2017, 391, 109–115. [Google Scholar] [CrossRef] [Green Version]
  36. Xu, H.; DeLuca, S.Z.; O’Farrell, P.H. Manipulating the Metazoan Mitochondrial Genome with Targeted Restriction Enzymes. Science 2008, 321, 575–577. [Google Scholar] [CrossRef] [Green Version]
  37. Silva-Pinheiro, P.; Minczuk, M. The Potential of Mitochondrial Genome Engineering. Nat. Rev. Genet. 2022, 23, 199–214. [Google Scholar] [CrossRef]
  38. Cho, J.; Hur, J.H.; Graniel, J.; Benzer, S.; Walker, D.W. Expression of Yeast NDI1 Rescues a Drosophila Complex I Assembly Defect. PLoS ONE 2012, 7, e50644. [Google Scholar] [CrossRef] [Green Version]
  39. Ghezzi, D.; Goffrini, P.; Uziel, G.; Horvath, R.; Klopstock, T.; Lochmüller, H.; D’Adamo, P.; Gasparini, P.; Strom, T.M.; Prokisch, H.; et al. SDHAF1, Encoding a LYR Complex-II Specific Assembly Factor, Is Mutated in SDH-Defective Infantile Leukoencephalopathy. Nat. Genet. 2009, 41, 654–656. [Google Scholar] [CrossRef]
  40. Hao, H.X.; Khalimonchuk, O.; Schraders, M.; Dephoure, N.; Bayley, J.P.; Kunst, H.; Devilee, P.; Cremers, C.W.R.J.; Schiffman, J.D.; Bentz, B.G.; et al. SDH5, a Gene Required for Flavination of Succinate Dehydrogenase, Is Mutated in Paraganglioma. Science 2009, 325, 1139–1142. [Google Scholar] [CrossRef] [Green Version]
  41. Mast, J.D.; Tomalty, K.M.H.; Vogel, H.; Clandinin, T.R. Reactive Oxygen Species Act Remotely to Cause Synapese Loss in a Drosophila Model of Developmental Mitochondrial Encephalopathy. Development 2008, 135, 2669–2679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Fan, F.; Sam, R.; Ryan, E.; Alvarado, K.; Villa-Cuesta, E. Rapamycin as a Potential Treatment for Succinate Dehydrogenase Deficiency. Heliyon 2019, 5, e01217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Walker, D.W.; Hájek, P.; Muffat, J.; Knoepfle, D.; Cornelison, S.; Attardi, G.; Benzer, S. Hypersensitivity to Oxygen and Shortened Lifespan in a Drosophila Mitochondrial Complex II Mutant. Proc. Natl. Acad. Sci. USA 2006, 103, 16382–16387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Tsuda, M.; Sugiura, T.; Ishii, T.; Ishii, N.; Aigaki, T. A Mev-1-like Dominant-Negative SdhC Increases Oxidative Stress and Reduces Lifespan in Drosophila. Biochem. Biophys. Res. Commun. 2007, 363, 342–346. [Google Scholar] [CrossRef]
  45. Van Vranken, J.G.; Bricker, D.K.; Dephoure, N.; Gygi, S.P.; Cox, J.E.; Thummel, C.S.; Rutter, J. SDHAF4 Promotes Mitochondrial Succinate Dehydrogenase Activity and Prevents Neurodegeneration. Cell Metab. 2014, 20, 241–252. [Google Scholar] [CrossRef] [Green Version]
  46. Na, U.; Yu, W.; Cox, J.; Bricker, D.K.; Brockmann, K.; Rutter, J.; Thummel, C.S.; Winge, D.R. The LYR Factors Sdhaf1 and SDHAF3 Mediate Maturation of the Iron-Sulfur Subunit of Succinate Dehydrogenase. Cell Metab. 2014, 20, 253–266. [Google Scholar] [CrossRef] [Green Version]
  47. Bayley, J.P.; Devilee, P. Hypothesis: Why Different Types of SDH Gene Variants Cause Divergent Tumor Phenotypes. Genes 2022, 13, 1025. [Google Scholar] [CrossRef]
  48. Fernández-Vizarra, E.; Zeviani, M. Nuclear Gene Mutations as the Cause of Mitochondrial Complex III Deficiency. Front. Genet. 2015, 6, 134. [Google Scholar] [CrossRef]
  49. Ghezzi, D.; Arzuffi, P.; Zordan, M.; Da Re, C.; Lamperti, C.; Benna, C.; D’Adamo, P.; Diodato, D.; Costa, R.; Mariotti, C.; et al. Mutations in TTC19 Cause Mitochondrial Complex III Deficiency and Neurological Impairment in Humans and Flies. Nat. Genet. 2011, 43, 259–263. [Google Scholar] [CrossRef]
  50. Peruzzo, R.; Corrà, S.; Costa, R.; Brischigliaro, M.; Varanita, T.; Biasutto, L.; Rampazzo, C.; Ghezzi, D.; Leanza, L.; Zoratti, M.; et al. Exploiting Pyocyanin to Treat Mitochondrial Disease Due to Respiratory Complex III Dysfunction. Nat. Commun. 2021, 12, 2103. [Google Scholar] [CrossRef]
  51. Brischigliaro, M.; Frigo, E.; Corrà, S.; De Pittà, C.; Szabò, I.; Zeviani, M.; Costa, R. Modelling of BCS1L-Related Human Mitochondrial Disease in Drosophila Melanogaster. J. Mol. Med. 2021, 99, 1471–1485. [Google Scholar] [CrossRef] [PubMed]
  52. Frolov, M.V.; Benevolenskaya, E.V.; Birchler, J.A. The Oxen Gene of Drosophila Encodes a Homolog of Subunit 9 of Yeast Ubiquinol-Cytochrome c Oxidoreductase Complex: Evidence for Modulation of Gene Expression in Response to Mitochondrial Activity. Genetics 2000, 156, 1727–1736. [Google Scholar] [CrossRef] [PubMed]
  53. Visapää, I.; Fellman, V.; Vesa, J.; Dasvarma, A.; Hutton, J.L.; Kumar, V.; Payne, G.S.; Makarow, M.; Van Coster, R.; Taylor, R.W.; et al. GRACILE Syndrome, a Lethal Metabolic Disorder with Iron Overload, Is Caused by a Point Mutation in BCS1L. Am. J. Hum. Genet. 2002, 71, 863–876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Baker, R.A.; Priestley, J.R.C.; Wilstermann, A.M.; Reese, K.J.; Mark, P.R. Clinical Spectrum of BCS1L Mitopathies and Their Underlying Structural Relationships. Am. J. Med. Genet. Part A 2019, 179, 373–380. [Google Scholar] [CrossRef]
  55. Calvo, S.E.; Compton, A.G.; Hershman, S.G.; Lim, S.C.; Lieber, D.S.; Tucker, E.J.; Laskowski, A.; Garone, C.; Liu, S.; Jaffe, D.B.; et al. Molecular Diagnosis of Infantile Mitochondrial Disease with Targeted Next-Generation Sequencing. Sci. Transl. Med. 2012, 4, 118ra10. [Google Scholar] [CrossRef] [Green Version]
  56. Bosch, J.A.; Ugur, B.; Pichardo-Casas, I.; Rabasco, J.; Escobedo, F.; Zuo, Z.; Brown, B.; Celniker, S.; Sinclair, D.A.; Bellen, H.J.; et al. Two Neuronal Peptides Encoded from a Single Transcript Regulate Mitochondrial Complex III in Drosophila. eLife 2022, 11, e82709. [Google Scholar] [CrossRef]
  57. Zhang, S.; Reljić, B.; Liang, C.; Kerouanton, B.; Francisco, J.C.; Peh, J.H.; Mary, C.; Jagannathan, N.S.; Olexiouk, V.; Tang, C.; et al. Mitochondrial Peptide BRAWNIN Is Essential for Vertebrate Respiratory Complex III Assembly. Nat. Commun. 2020, 11, 1312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Dennerlein, S.; Poerschke, S.; Oeljeklaus, S.; Wang, C.; Richter-Dennerlein, R.; Sattmann, J.; Bauermeister, D.; Hanitsch, E.; Stoldt, S.; Langer, T.; et al. Defining the Interactome of the Human Mitochondrial Ribosome Identifies SMIM4 and TMEM223 as Respiratory Chain Assembly Factors. eLife 2021, 10, e68213. [Google Scholar] [CrossRef]
  59. Brischigliaro, M.; Zeviani, M. Cytochrome c Oxidase Deficiency. Biochim. Biophys. Acta Bioenerg. 2021, 1862, 148335. [Google Scholar] [CrossRef]
  60. Massa, V.; Fernandez-Vizarra, E.; Alshahwan, S.; Bakhsh, E.; Goffrini, P.; Ferrero, I.; Mereghetti, P.; D’Adamo, P.; Gasparini, P.; Zeviani, M. Severe Infantile Encephalomyopathy Caused by a Mutation in COX6B1, a Nucleus-Encoded Subunit of Cytochrome C Oxidase. Am. J. Hum. Genet. 2008, 82, 1281–1289. [Google Scholar] [CrossRef] [Green Version]
  61. Shoubridge, E.A. Cytochrome c Oxidase Deficiency. Am. J. Med. Genet. Semin. Med. Genet. 2001, 106, 46–52. [Google Scholar] [CrossRef] [PubMed]
  62. Darin, N.; Moslemi, A.R.; Lebon, S.; Rustin, P.; Holme, E.; Oldfors, A.; Tulinius, M. Genotypes and Clinical Phenotypes in Children with Cytochrome-c Oxidase Deficiency. Neuropediatrics 2003, 34, 311–317. [Google Scholar] [CrossRef] [PubMed]
  63. Timón-Gómez, A.; Nývltová, E.; Abriata, L.A.; Vila, A.J.; Hosler, J.; Barrientos, A. Mitochondrial Cytochrome c Oxidase Biogenesis: Recent Developments. Semin. Cell Dev. Biol. 2018, 76, 163–178. [Google Scholar] [CrossRef] [PubMed]
  64. Brischigliaro, M.; Cabrera-Orefice, A.; Sturlese, M.; Elurbe, D.M.; Frigo, E.; Fernandez-Vizarra, E.; Moro, S.; Huynen, M.A.; Arnold, S.; Viscomi, C.; et al. CG7630 Is the Drosophila Melanogaster Homolog of the Cytochrome c Oxidase Subunit COX7B. EMBO Rep. 2022, 23, e54825. [Google Scholar] [CrossRef] [PubMed]
  65. Patel, M.R.; Miriyala, G.K.; Littleton, A.J.; Yang, H.; Trinh, K.; Young, J.M.; Kennedy, S.R.; Yamashita, Y.M.; Pallanck, L.J.; Malik, H.S. A Mitochondrial DNA Hypomorph of Cytochrome Oxidase Specifically Impairs Male Fertility in Drosophila Melanogaster. eLife 2016, 5, e16923. [Google Scholar] [CrossRef]
  66. Szuplewski, S.; Terracol, R. The Cyclope Gene of Drosophila Encodes a Cytochrome c Oxidase Subunit VIc Homolog. Genetics 2001, 158, 1629–1643. [Google Scholar] [CrossRef]
  67. Mandal, S.; Guptan, P.; Owusu-Ansah, E.; Banerjee, U. Mitochondrial Regulation of Cell Cycle Progression during Development as Revealed by the Tenured Mutation in Drosophila. Dev. Cell 2005, 9, 843–854. [Google Scholar] [CrossRef] [Green Version]
  68. Liu, W.; Gnanasambandam, R.; Benjamin, J.; Kaur, G.; Getman, P.B.; Siegel, A.J.; Shortridge, R.D.; Singh, S. Mutations in Cytochrome c Oxidase Subunit VIa Cause Neurodegeneration and Motor Dysfunction in Drosophila. Genetics 2007, 176, 937–946. [Google Scholar] [CrossRef] [Green Version]
  69. Klichko, V.; Sohal, B.H.; Radyuk, S.N.; Orr, W.C.; Sohal, R.S. Decrease in Cytochrome c Oxidase Reserve Capacity Diminishes Robustness of Drosophila Melanogaster and Shortens Lifespan. Biochem. J. 2014, 459, 127–135. [Google Scholar] [CrossRef]
  70. Kemppainen, K.K.; Rinne, J.; Sriram, A.; Lakanmaa, M.; Zeb, A.; Tuomela, T.; Popplestone, A.; Singh, S.; Sanz, A.; Rustin, P.; et al. Expression of Alternative Oxidase in Drosophila Ameliorates Diverse Phenotypes Due to Cytochrome Oxidase Deficiency. Hum. Mol. Genet. 2014, 23, 2078–2093. [Google Scholar] [CrossRef] [Green Version]
  71. Peralta, S.; Clemente, P.; Sánchez-Martínez, Á.; Calleja, M.; Hernández-Sierra, R.; Matsushima, Y.; Adán, C.; Ugalde, C.; Fernández-Moreno, M.Á.; Kaguni, L.S.; et al. Coiled Coil Domain-Containing Protein 56 (CCDC56) Is a Novel Mitochondrial Protein Essential for Cytochrome c Oxidase Function. J. Biol. Chem. 2012, 287, 24174–24185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Porcelli, D.; Oliva, M.; Duchi, S.; Latorre, D.; Cavaliere, V.; Barsanti, P.; Villani, G.; Gargiulo, G.; Caggese, C. Genetic, Functional and Evolutionary Characterization of Scox, the Drosophila Melanogaster Ortholog of the Human SCO1 Gene. Mitochondrion 2010, 10, 433–448. [Google Scholar] [CrossRef] [PubMed]
  73. Nguyen, T.B.; Ida, H.; Shimamura, M.; Kitazawa, D.; Akao, S.; Yoshida, H.; Inoue, Y.H.; Yamaguchi, M. Role of SCOX in Determination of Drosophila Melanogaster Lifespan. Am. J. Cancer Res. 2014, 4, 325–336. [Google Scholar] [PubMed]
  74. Martínez-Morentin, L.; Martínez, L.; Piloto, S.; Yang, H.; Schon, E.A.; Garesse, R.; Bodmer, R.; Ocorr, K.; Cervera, M.; Arredondo, J.J. Cardiac Deficiency of Single Cytochrome Oxidase Assembly Factor Scox Induces P53-Dependent Apoptosis in a Drosophila Cardiomyopathy Model. Hum. Mol. Genet. 2015, 24, 3608–3622. [Google Scholar] [CrossRef] [Green Version]
  75. Zordan, M.A.; Cisotto, P.; Benna, C.; Agostino, A.; Rizzo, G.; Piccin, A.; Pegoraro, M.; Sandrelli, F.; Perini, G.; Tognon, G.; et al. Post-Transcriptional Silencing and Functional Characterization of the Drosophila Melanogaster Homolog of Human Surf1. Genetics 2006, 172, 229–241. [Google Scholar] [CrossRef] [Green Version]
  76. Da-Rè, C.; von Stockum, S.; Biscontin, A.; Millino, C.; Cisotto, P.; Zordan, M.A.; Zeviani, M.; Bernardi, P.; De Pittà, C.; Costa, R. Leigh Syndrome in Drosophila Melanogaster. J. Biol. Chem. 2014, 289, 29235–29246. [Google Scholar] [CrossRef] [Green Version]
  77. Higuchi, Y.; Okunushi, R.; Hara, T.; Hashiguchi, A.; Yuan, J.; Yoshimura, A.; Murayama, K.; Ohtake, A.; Ando, M.; Hiramatsu, Y.; et al. Mutations in COA7 Cause Spinocerebellar Ataxia with Axonal Neuropathy. Brain 2018, 141, 1622–1636. [Google Scholar] [CrossRef]
  78. Brischigliaro, M.; Corrà, S.; Tregnago, C.; Fernandez-Vizarra, E.; Zeviani, M.; Costa, R.; De Pittà, C. Knockdown of APOPT1/COA8 Causes Cytochrome c Oxidase Deficiency, Neuromuscular Impairment, and Reduced Resistance to Oxidative Stress in Drosophila Melanogaster. Front. Physiol. 2019, 10, 1143. [Google Scholar] [CrossRef]
  79. Ostergaard, E.; Weraarpachai, W.; Ravn, K.; Born, A.P.; Jønson, L.; Duno, M.; Wibrand, F.; Shoubridge, E.A.; Vissing, J. Mutations in COA3 Cause Isolated Complex IV Deficiency Associated with Neuropathy, Exercise Intolerance, Obesity, and Short Stature. J. Med. Genet. 2015, 52, 203–207. [Google Scholar] [CrossRef]
  80. Kowada, R.; Kodani, A.; Ida, H.; Yamaguchi, M.; Lee, I.S.; Okada, Y.; Yoshida, H. The Function of Scox in Glial Cells Is Essential for Locomotive Ability in Drosophila. Sci. Rep. 2021, 11, 21207. [Google Scholar] [CrossRef]
  81. Brischigliaro, M.; Badocco, D.; Costa, R.; Viscomi, C.; Zeviani, M.; Pastore, P.; Fernández-Vizarra, E. Mitochondrial Cytochrome c Oxidase Defects Alter Cellular Homeostasis of Transition Metals. Front. Cell Dev. Biol. 2022, 10, 1090. [Google Scholar] [CrossRef]
  82. Sánchez-Caballero, L.; Elurbe, D.M.; Baertling, F.; Guerrero-Castillo, S.; van den Brand, M.; van Strien, J.; van Dam, T.J.P.; Rodenburg, R.; Brandt, U.; Huynen, M.A.; et al. TMEM70 Functions in the Assembly of Complexes I and V. Biochim. Biophys. Acta Bioenerg. 2020, 1861, 148202. [Google Scholar] [CrossRef]
  83. Carroll, J.; He, J.; Ding, S.; Fearnley, I.M.; Walker, J.E. TMEM70 and TMEM242 Help to Assemble the Rotor Ring of Human ATP Synthase and Interact with Assembly Factors for Complex I. Proc. Natl. Acad. Sci. USA 2021, 118, e2100558118. [Google Scholar] [CrossRef] [PubMed]
  84. Celotto, A.M.; Frank, A.C.; McGrath, S.W.; Fergestad, T.; Van Voorhies, W.A.; Buttle, K.F.; Mannella, C.A.; Palladino, M.J. Mitochondrial Encephalomyopathy in Drosophila. J. Neurosci. 2006, 26, 810–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Sun, X.; Wheeler, C.T.; Yolitz, J.; Laslo, M.; Alberico, T.; Sun, Y.; Song, Q.; Zou, S. A Mitochondrial ATP Synthase Subunit Interacts with TOR Signaling to Modulate Protein Homeostasis and Lifespan in Drosophila. Cell Rep. 2014, 8, 1781–1792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Lovero, D.; Giordano, L.; Marsano, R.M.; Sanchez-Martinez, A.; Boukhatmi, H.; Drechsler, M.; Oliva, M.; Whitworth, A.J.; Porcelli, D.; Caggese, C. Characterization of Drosophila ATPsynC Mutants as a New Model of Mitochondrial ATP Synthase Disorders. PLoS ONE 2018, 13, e0201811. [Google Scholar] [CrossRef] [Green Version]
  87. Chen, Y.N.; Wu, C.H.; Zheng, Y.; Li, J.J.; Wang, J.L.; Wang, Y.F. Knockdown of ATPsyn-b Caused Larval Growth Defect and Male Infertility in Drosophila. Arch. Insect Biochem. Physiol. 2015, 88, 144–154. [Google Scholar] [CrossRef]
  88. Teixeira, F.K.; Sanchez, C.G.; Hurd, T.R.; Seifert, J.R.K.; Czech, B.; Preall, J.B.; Hannon, G.J.; Lehmann, R. ATP Synthase Promotes Germ Cell Differentiation Independent of Oxidative Phosphorylation. Nat. Cell Biol. 2015, 17, 689–696. [Google Scholar] [CrossRef] [Green Version]
  89. Alcázar-Fabra, M.; Rodríguez-Sánchez, F.; Trevisson, E.; Brea-Calvo, G. Primary Coenzyme Q Deficiencies: A Literature Review and Online Platform of Clinical Features to Uncover Genotype-Phenotype Correlations. Free Radic. Biol. Med. 2021, 167, 141–180. [Google Scholar] [CrossRef]
  90. Grant, J.; Saldanha, J.W.; Gould, A.P. A Drosophila Model for Primary Coenzyme Q Deficiency and Dietary Rescue in the Developing Nervous System. DMM Dis. Model. Mech. 2010, 3, 799–806. [Google Scholar] [CrossRef] [Green Version]
  91. Almannai, M.; El-Hattab, A.W.; Scaglia, F. Mitochondrial DNA Replication: Clinical Syndromes. Essays Biochem. 2018, 62, 297–308. [Google Scholar] [CrossRef] [PubMed]
  92. Chapman, J.; Ng, Y.S.; Nicholls, T.J. The Maintenance of Mitochondrial DNA Integrity and Dynamics by Mitochondrial Membranes. Life 2020, 10, 164. [Google Scholar] [CrossRef] [PubMed]
  93. Iyengar, B.; Roote, J.; Campos, A.R. The Tamas Gene, Identified as a Mutation That Disrupts Larval Behavior in Drosophila Melanogaster, Codes for the Mitochondrial DNA Polymerase Catalytic Subunit (DNApol-Γ125). Genetics 1999, 153, 1809–1824. [Google Scholar] [CrossRef]
  94. Bratic, A.; Kauppila, T.E.S.; Macao, B.; Grönke, S.; Siibak, T.; Stewart, J.B.; Baggio, F.; Dols, J.; Partridge, L.; Falkenberg, M.; et al. Complementation between Polymerase- and Exonuclease-Deficient Mitochondrial DNA Polymerase Mutants in Genomically Engineered Flies. Nat. Commun. 2015, 6, 8808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Rodrigues, A.P.C.; Camargo, A.F.; Andjelković, A.; Jacobs, H.T.; Oliveira, M.T. Developmental Arrest in Drosophila Melanogaster Caused by Mitochondrial DNA Replication Defects Cannot Be Rescued by the Alternative Oxidase. Sci. Rep. 2018, 8, 10882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Samstag, C.L.; Hoekstra, J.G.; Huang, C.H.; Chaisson, M.J.; Youle, R.J.; Kennedy, S.R.; Pallanck, L.J. Deleterious Mitochondrial DNA Point Mutations Are Overrepresented in Drosophila Expressing a Proofreading-Defective DNA Polymerase γ. PLoS Genet. 2018, 14, e1007805. [Google Scholar] [CrossRef] [PubMed]
  97. Vermulst, M.; Bielas, J.H.; Kujoth, G.C.; Ladiges, W.C.; Rabinovitch, P.S.; Prolla, T.A.; Loeb, L.A. Mitochondrial Point Mutations Do Not Limit the Natural Lifespan of Mice. Nat. Genet. 2007, 39, 540–543. [Google Scholar] [CrossRef]
  98. Vermulst, M.; Wanagat, J.; Kujoth, G.C.; Bielas, J.H.; Rabinovitch, P.S.; Prolla, T.A.; Loeb, L.A. DNA Deletions and Clonal Mutations Drive Premature Aging in Mitochondrial Mutator Mice. Nat. Genet. 2008, 40, 392–394. [Google Scholar] [CrossRef]
  99. Edgar, D.; Shabalina, I.; Camara, Y.; Wredenberg, A.; Calvaruso, M.A.; Nijtmans, L.; Nedergaard, J.; Cannon, B.; Larsson, N.G.; Trifunovic, A. Random Point Mutations with Major Effects on Protein-Coding Genes Are the Driving Force behind Premature Aging in MtDNA Mutator Mice. Cell Metab. 2009, 10, 131–138. [Google Scholar] [CrossRef] [Green Version]
  100. Andreazza, S.; Samstag, C.L.; Sanchez-Martinez, A.; Fernandez-Vizarra, E.; Gomez-Duran, A.; Lee, J.J.; Tufi, R.; Hipp, M.J.; Schmidt, E.K.; Nicholls, T.J.; et al. Mitochondrially-Targeted APOBEC1 Is a Potent MtDNA Mutator Affecting Mitochondrial Function and Organismal Fitness in Drosophila. Nat. Commun. 2019, 10, 3280. [Google Scholar] [CrossRef] [Green Version]
  101. Sanchez-Martinez, A.; Calleja, M.; Peralta, S.; Matsushima, Y.; Hernandez-Sierra, R.; Whitworth, A.J.; Kaguni, L.S.; Garesse, R. Modeling Pathogenic Mutations of Human Twinkle in Drosophila Suggests an Apoptosis Role in Response to Mitochondrial Defects. PLoS ONE 2012, 7, e43954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. He, J.; Cooper, H.M.; Reyes, A.; Di Re, M.; Sembongi, H.; Gao, J.; Neuman, K.C.; Fearnley, I.M.; Spinazzola, A.; Walker, J.E.; et al. Mitochondrial Nucleoid Interacting Proteins Support Mitochondrial Protein Synthesis. Nucleic Acids Res. 2012, 40, 6109–6121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Harel, T.; Yoon, W.H.; Garone, C.; Gu, S.; Coban-Akdemir, Z.; Eldomery, M.K.; Posey, J.E.; Jhangiani, S.N.; Rosenfeld, J.A.; Cho, M.T.; et al. Recurrent De Novo and Biallelic Variation of ATAD3A, Encoding a Mitochondrial Membrane Protein, Results in Distinct Neurological Syndromes. Am. J. Hum. Genet. 2016, 99, 831–845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Frazier, A.E.; Compton, A.G.; Kishita, Y.; Hock, D.H.; Welch, A.M.E.; Amarasekera, S.S.C.; Rius, R.; Formosa, L.E.; Imai-Okazaki, A.; Francis, D.; et al. Fatal Perinatal Mitochondrial Cardiac Failure Caused by Recurrent De Novo Duplications in the ATAD3 Locus. Med 2021, 2, 49–73.e10. [Google Scholar] [CrossRef] [PubMed]
  105. Yap, Z.Y.; Park, Y.H.; Wortmann, S.B.; Gunning, A.C.; Ezer, S.; Lee, S.; Duraine, L.; Wilichowski, E.; Wilson, K.; Mayr, J.A.; et al. Functional Interpretation of ATAD3A Variants in Neuro-Mitochondrial Phenotypes. Genome Med. 2021, 13, 55. [Google Scholar] [CrossRef]
  106. Gilquin, B.; Taillebourg, E.; Cherradi, N.; Hubstenberger, A.; Gay, O.; Merle, N.; Assard, N.; Fauvarque, M.-O.; Tomohiro, S.; Kuge, O.; et al. The AAA + ATPase ATAD3A Controls Mitochondrial Dynamics at the Interface of the Inner and Outer Membranes. Mol. Cell. Biol. 2010, 30, 1984–1996. [Google Scholar] [CrossRef] [Green Version]
  107. Ostergaard, E.; Christensen, E.; Kristensen, E.; Mogensen, B.; Duno, M.; Shoubridge, E.A.; Wibrand, F. Deficiency of the α Subunit of Succinate-Coenzyme A Ligase Causes Fatal Infantile Lactic Acidosis with Mitochondrial DNA Depletion. Am. J. Hum. Genet. 2007, 81, 383–387. [Google Scholar] [CrossRef] [Green Version]
  108. Carrozzo, R.; Verrigni, D.; Rasmussen, M.; de Coo, R.; Amartino, H.; Bianchi, M.; Buhas, D.; Mesli, S.; Naess, K.; Born, A.P.; et al. Succinate-CoA Ligase Deficiency Due to Mutations in SUCLA2 and SUCLG1: Phenotype and Genotype Correlations in 71 Patients. J. Inherit. Metab. Dis. 2016, 39, 243–252. [Google Scholar] [CrossRef]
  109. Demirbas, D.; Harris, D.J.; Arn, P.H.; Huang, X.; Waisbren, S.E.; Anselm, I.; Lerner-Ellis, J.P.; Wong, L.J.; Levy, H.L.; Berry, G.T. Phenotypic Variability in Deficiency of the α Subunit of Succinate-CoA Ligase. JIMD Rep. 2019, 46, 63–69. [Google Scholar] [CrossRef]
  110. Quan, X.; Sato-Miyata, Y.; Tsuda, M.; Muramatsu, K.; Asano, T.; Takeo, S.; Aigaki, T. Deficiency of Succinyl-CoA Synthetase α Subunit Delays Development, Impairs Locomotor Activity and Reduces Survival under Starvation in Drosophila. Biochem. Biophys. Res. Commun. 2017, 483, 566–571. [Google Scholar] [CrossRef]
  111. Kodani, A.; Yamaguchi, M.; Itoh, R.; Huynh, M.A.; Yoshida, H. A Drosophila Model of the Neurological Symptoms in Mpv17-Related Diseases. Sci. Rep. 2022, 12, 22632. [Google Scholar] [CrossRef] [PubMed]
  112. Spinazzola, A.; Viscomi, C.; Fernandez-Vizarra, E.; Carrara, F.; D’Adamo, P.; Calvo, S.; Marsano, R.M.; Donnini, C.; Weiher, H.; Strisciuglio, P.; et al. MPV17 Encodes an Inner Mitochondrial Membrane Protein and Is Mutated in Infantile Hepatic Mitochondrial DNA Depletion. Nat. Genet. 2006, 38, 570–576. [Google Scholar] [CrossRef] [PubMed]
  113. Mendelsohn, B.A.; Mehta, N.; Hameed, B.; Pekmezci, M.; Packman, S.; Ralph, J. Adult-Onset Fatal Neurohepatopathy in a Woman Caused by MPV17 Mutation. JIMD Rep. 2014, 13, 37–41. [Google Scholar] [CrossRef] [Green Version]
  114. D’Souza, A.R.; Minczuk, M. Mitochondrial Transcription and Translation: Overview. Essays Biochem. 2018, 62, 309–320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Boczonadi, V.; Horvath, R. Mitochondria: Impaired Mitochondrial Translation in Human Disease. Int. J. Biochem. Cell Biol. 2014, 48, 77–84. [Google Scholar] [CrossRef] [Green Version]
  116. Ferrari, A.; Del’Olio, S.; Barrientos, A. The Diseased Mitoribosome. FEBS Lett. 2021, 595, 1025–1061. [Google Scholar] [CrossRef]
  117. Royden, C.S.; Pirrotta, V.; Jan, L.Y. The Tko Locus, Site of a Behavioral Mutation in D. Melanogaster, Codes for a Protein Homologous to Prokaryotic Ribosomal Protein S12. Cell 1987, 51, 165–173. [Google Scholar] [CrossRef]
  118. Toivonen, J.M.; O’Dell, K.M.C.; Petit, N.; Irvine, S.C.; Knight, G.K.; Lehtonen, M.; Longmuir, M.; Luoto, K.; Touraille, S.; Wang, Z.; et al. Technical Knockout, a Drosophila Model of Mitochondrial Deafness. Genetics 2001, 159, 241–254. [Google Scholar] [CrossRef]
  119. Gokhale, A.; Lee, C.E.; Zlatic, S.A.; Freeman, A.A.H.; Shearing, N.; Hartwig, C.; Ogunbona, O.; Bassell, J.L.; Wynne, M.E.; Werner, E.; et al. Mitochondrial Proteostasis Requires Genes Encoded in a Neurodevelopmental Syndrome Locus. J. Neurosci. 2021, 41, 6596–6616. [Google Scholar] [CrossRef]
  120. Tilokani, L.; Nagashima, S.; Paupe, V.; Prudent, J. Mitochondrial Dynamics: Overview of Molecular Mechanisms. Essays Biochem. 2018, 62, 341–360. [Google Scholar] [CrossRef] [Green Version]
  121. Hales, K.G.; Fuller, M.T. Developmentally Regulated Mitochondrial Fusion Mediated by a Conserved, Novel, Predicted GTPase. Cell 1997, 90, 121–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Chen, H.; Detmer, S.A.; Ewald, A.J.; Griffin, E.E.; Fraser, S.E.; Chan, D.C. Mitofusins Mfn1 and Mfn2 Coordinately Regulate Mitochondrial Fusion and Are Essential for Embryonic Development. J. Cell Biol. 2003, 160, 189–200. [Google Scholar] [CrossRef] [PubMed]
  123. Ishihara, N.; Nomura, M.; Jofuku, A.; Kato, H.; Suzuki, S.O.; Masuda, K.; Otera, H.; Nakanishi, Y.; Nonaka, I.; Goto, Y.I.; et al. Mitochondrial Fission Factor Drp1 Is Essential for Embryonic Development and Synapse Formation in Mice. Nat. Cell Biol. 2009, 11, 958–966. [Google Scholar] [CrossRef] [PubMed]
  124. Wakabayashi, J.; Zhang, Z.; Wakabayashi, N.; Tamura, Y.; Fukaya, M.; Kensler, T.W.; Iijima, M.; Sesaki, H. The Dynamin-Related GTPase Drp1 Is Required for Embryonic and Brain Development in Mice. J. Cell Biol. 2009, 186, 805–816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Chan, D.C. Mitochondrial Dynamics and Its Involvement in Disease. Annu. Rev. Pathol. Mech. Dis. 2020, 15, 235–259. [Google Scholar] [CrossRef] [Green Version]
  126. Verstreken, P.; Ly, C.V.; Venken, K.J.T.; Koh, T.W.; Zhou, Y.; Bellen, H.J. Synaptic Mitochondria Are Critical for Mobilization of Reserve Pool Vesicles at Drosophila Neuromuscular Junctions. Neuron 2005, 47, 365–378. [Google Scholar] [CrossRef] [Green Version]
  127. Aldridge, A.C.; Benson, L.P.; Siegenthaler, M.M.; Whigham, B.T.; Stowers, R.S.; Hales, K.G. Roles for Drp1, a Dynamin-Related Protein, and Milton, a Kinesin-Associated Protein, in Mitochondrial Segregation, Unfurling and Elongation during Drosophila Spermatogenesis. Fly 2007, 1, 38–46. [Google Scholar] [CrossRef] [Green Version]
  128. Yarosh, W.; Monserrate, J.; Tong, J.J.; Tse, S.; Le, P.K.; Nguyen, K.; Brachmann, C.B.; Wallace, D.C.; Huang, T. The Molecular Mechanisms of OPA1-Mediated Optic Atrophy in Drosophila Model and Prospects for Antioxidant Treatment. PLoS Genet. 2008, 4, 0062–0071. [Google Scholar] [CrossRef] [Green Version]
  129. Shahrestani, P.; Leung, H.T.; Le, P.K.; Pak, W.L.; Tse, S.; Ocorr, K.; Huang, T. Heterozygous Mutation of Drosophila Opa1 Causes the Development of Multiple Organ Abnormalities in an Age-Dependent and Organ-Specific Manner. PLoS ONE 2009, 4, e6867. [Google Scholar] [CrossRef] [Green Version]
  130. Tang, S.; Le, P.K.; Tse, S.; Wallace, D.C.; Huang, T. Heterozygous Mutation of Opa1 in Drosophila Shortens Lifespan Mediated through Increased Reactive Oxygen Species Production. PLoS ONE 2009, 4, e4492. [Google Scholar] [CrossRef] [Green Version]
  131. Dorn, G.W.; Clark, C.F.; Eschenbacher, W.H.; Kang, M.Y.; Engelhard, J.T.; Warner, S.J.; Matkovich, S.J.; Jowdy, C.C. MARF and Opa1 Control Mitochondrial and Cardiac Function in Drosophila. Circ. Res. 2011, 108, 12–17. [Google Scholar] [CrossRef] [Green Version]
  132. Davies, K.M.; Anselmi, C.; Wittig, I.; Faraldo-Gómez, J.D.; Kühlbrandt, W. Structure of the Yeast F 1F O-ATP Synthase Dimer and Its Role in Shaping the Mitochondrial Cristae. Proc. Natl. Acad. Sci. USA 2012, 109, 13602–13607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Blum, T.B.; Hahn, A.; Meier, T.; Davies, K.M.; Kühlbrandt, W. Dimers of Mitochondrial ATP Synthase Induce Membrane Curvature and Self-Assemble into Rows. Proc. Natl. Acad. Sci. USA 2019, 116, 4250–4255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Mukherjee, I.; Ghosh, M.; Meinecke, M. MICOS and the Mitochondrial Inner Membrane Morphology—When Things Get out of Shape. FEBS Lett. 2021, 595, 1159–1183. [Google Scholar] [CrossRef] [PubMed]
  135. Guarani, V.; Jardel, C.; Chrétien, D.; Lombès, A.; Bénit, P.; Labasse, C.; Lacène, E.; Bourillon, A.; Imbard, A.; Benoist, J.F.; et al. QIL1 Mutation Causes MICOS Disassembly and Early Onset Fatal Mitochondrial Encephalopathy with Liver Disease. eLife 2016, 5, e17163. [Google Scholar] [CrossRef]
  136. Zeharia, A.; Friedman, J.R.; Tobar, A.; Saada, A.; Konen, O.; Fellig, Y.; Shaag, A.; Nunnari, J.; Elpeleg, O. Mitochondrial Hepato-Encephalopathy Due to Deficiency of QIL1/MIC13 (C19orf70), a MICOS Complex Subunit. Eur. J. Hum. Genet. 2016, 24, 1778–1782. [Google Scholar] [CrossRef] [Green Version]
  137. Russell, B.E.; Whaley, K.G.; Bove, K.E.; Labilloy, A.; Lombardo, R.C.; Hopkin, R.J.; Leslie, N.D.; Prada, C.; Assouline, Z.; Barcia, G.; et al. Expanding and Underscoring the Hepato-Encephalopathic Phenotype of QIL1/MIC13. Hepatology 2019, 70, 1066–1070. [Google Scholar] [CrossRef]
  138. Benincá, C.; Zanette, V.; Brischigliaro, M.; Johnson, M.; Reyes, A.; Valle, D.A.D.; Robinson, A.J.; Degiorgi, A.; Yeates, A.; Telles, B.A.; et al. Mutation in the MICOS Subunit Gene APOO (MIC26) Associated with an X-Linked Recessive Mitochondrial Myopathy, Lactic Acidosis, Cognitive Impairment and Autistic Features. J. Med. Genet. 2021, 58, 155–167. [Google Scholar] [CrossRef]
  139. 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]
  140. Tsai, P.I.; Papakyrikos, A.M.; Hsieh, C.H.; Wang, X. Drosophila MIC60/Mitoflin Conducts Dual Roles in Mitochondrial Motility and Crista Structure. Mol. Biol. Cell 2017, 28, 3471–3479. [Google Scholar] [CrossRef]
  141. Guarani, V.; McNeill, E.M.; Paulo, J.A.; Huttlin, E.L.; Fröhlich, F.; Gygi, S.P.; Vactor, D.V.; Wade Harper, J. QIL1 Is a Novel Mitochondrial Protein Required for MICOS Complex Stability and Cristae Morphology. eLife 2015, 4, e06265. [Google Scholar] [CrossRef] [PubMed]
  142. Charlesworth, G.; Balint, B.; Mencacci, N.E.; Carr, L.; Wood, N.W.; Bhatia, K.P. SLC25A46 Mutations Underlie Progressive Myoclonic Ataxia with Optic Atrophy and Neuropathy. Mov. Disord. Off. J. Mov. Disord. Soc. 2016, 31, 1249–1251. [Google Scholar] [CrossRef] [PubMed]
  143. Janer, A.; Prudent, J.; Paupe, V.; Fahiminiya, S.; Majewski, J.; Sgarioto, N.; Des Rosiers, C.; Forest, A.; Lin, Z.; Gingras, A.; et al. SLC 25A46 Is Required for Mitochondrial Lipid Homeostasis and Cristae Maintenance and Is Responsible for Leigh Syndrome. EMBO Mol. Med. 2016, 8, 1019–1038. [Google Scholar] [CrossRef] [PubMed]
  144. Wan, J.; Steffen, J.; Yourshaw, M.; Mamsa, H.; Andersen, E.; Rudnik-Schöneborn, S.; Pope, K.; Howell, K.B.; McLean, C.A.; Kornberg, A.J.; et al. Loss of Function of SLC25A46 Causes Lethal Congenital Pontocerebellar Hypoplasia. Brain 2016, 139, 2877–2890. [Google Scholar] [CrossRef] [Green Version]
  145. Nguyen, M.; Boesten, I.; Hellebrekers, D.M.E.I.; Mulder-den Hartog, N.M.; de Coo, I.F.M.; Smeets, H.J.M.; Gerards, M. Novel Pathogenic SLC25A46 Splice-Site Mutation Causes an Optic Atrophy Spectrum Disorder. Clin. Genet. 2017, 91, 121–125. [Google Scholar] [CrossRef]
  146. Ali, M.S.; Suda, K.; Kowada, R.; Ueoka, I.; Yoshida, H.; Yamaguchi, M. Neuron-Specific Knockdown of Solute Carrier Protein SLC25A46a Induces Locomotive Defects, an Abnormal Neuron Terminal Morphology, Learning Disability, and Shortened Lifespan. IBRO Rep. 2020, 8, 65–75. [Google Scholar] [CrossRef]
  147. Garone, C.; Viscomi, C. Towards a Therapy for Mitochondrial Disease: An Update. Biochem. Soc. Trans. 2018, 46, 1247–1261. [Google Scholar] [CrossRef] [Green Version]
Figure 1. D. melanogaster life cycle at 25 °C. After mating between adult female and male flies, fertilized eggs are laid and the embryo develops into a first instar larva in ~24 h. Afterwards, the larvae grow and go through two additional larval stages (second and third instars), each one lasting ~24 h. During the larval stages, D. melanogaster exhibits high glycolytic flux, lactate production, and a high rate of glycogen synthesis and triglyceride (TAG) accumulation, which are needed for the metamorphosis. At the end of the third instar (2–3 days), larvae pupate. During the pupal stage, metamorphosis occurs (3–5 days) and adult fly tissues form. At the end of the metamorphosis, eclosion from the puparium occurs and adult flies become fertile after ~24 h. Flies live for 60–90 days depending on the rearing conditions (i.e., temperature and diet composition).
Figure 1. D. melanogaster life cycle at 25 °C. After mating between adult female and male flies, fertilized eggs are laid and the embryo develops into a first instar larva in ~24 h. Afterwards, the larvae grow and go through two additional larval stages (second and third instars), each one lasting ~24 h. During the larval stages, D. melanogaster exhibits high glycolytic flux, lactate production, and a high rate of glycogen synthesis and triglyceride (TAG) accumulation, which are needed for the metamorphosis. At the end of the third instar (2–3 days), larvae pupate. During the pupal stage, metamorphosis occurs (3–5 days) and adult fly tissues form. At the end of the metamorphosis, eclosion from the puparium occurs and adult flies become fertile after ~24 h. Flies live for 60–90 days depending on the rearing conditions (i.e., temperature and diet composition).
Biomolecules 13 00378 g001
Figure 2. Schematic representation of the most widely applied tools for the generation of D. melanogaster models of human disease. (A) The GAL4/UAS system is based on crossing between a driver (GAL4) line, expressing the GAL4 transcriptional activator under the control of an endogenous D. melanogaster promoter, and a responder (UAS) line, expressing a construct of interest (transgene or inverted-repeat sequence for RNAi). The progeny carries both constructs and the GAL4 activator protein binds the UAS to drive the expression of the downstream construct of interest. (B) Flies carrying transposable elements generated through germline transformation (microinjection of an embryo) can be crossed with the transposase flies expressing the transposase enzyme that will mobilize the transposable elements in the transformed flies. Excision can be precise (rescue of the endogenous locus) or imprecise (generation of new alleles). (C) Chemical mutagenesis in D. melanogaster is achieved by treating males with mutagens (e.g., EMS), which introduce GC to AT transitions in the germline. (D) TALEN genome editing is based on co-injection into the embryo of two vectors carrying two TALEN constructs (left/right TAL effectors in fusion with FokI nuclease). (E) CRISPR/Cas editing is based on co-injection of one vector carrying gRNA constructs and one vector carrying a Cas nuclease. Abbreviations: UAS—upstream activating sequence, dsRNA—double-strand RNA, EMS—ethyl methanesulfonate, TALEN—transcription activator-like effector nuclease, CRISPR—clustered regularly interspaced short palindromic repeats, gRNA—guide RNA, Cas—CRISPR-associated protein.
Figure 2. Schematic representation of the most widely applied tools for the generation of D. melanogaster models of human disease. (A) The GAL4/UAS system is based on crossing between a driver (GAL4) line, expressing the GAL4 transcriptional activator under the control of an endogenous D. melanogaster promoter, and a responder (UAS) line, expressing a construct of interest (transgene or inverted-repeat sequence for RNAi). The progeny carries both constructs and the GAL4 activator protein binds the UAS to drive the expression of the downstream construct of interest. (B) Flies carrying transposable elements generated through germline transformation (microinjection of an embryo) can be crossed with the transposase flies expressing the transposase enzyme that will mobilize the transposable elements in the transformed flies. Excision can be precise (rescue of the endogenous locus) or imprecise (generation of new alleles). (C) Chemical mutagenesis in D. melanogaster is achieved by treating males with mutagens (e.g., EMS), which introduce GC to AT transitions in the germline. (D) TALEN genome editing is based on co-injection into the embryo of two vectors carrying two TALEN constructs (left/right TAL effectors in fusion with FokI nuclease). (E) CRISPR/Cas editing is based on co-injection of one vector carrying gRNA constructs and one vector carrying a Cas nuclease. Abbreviations: UAS—upstream activating sequence, dsRNA—double-strand RNA, EMS—ethyl methanesulfonate, TALEN—transcription activator-like effector nuclease, CRISPR—clustered regularly interspaced short palindromic repeats, gRNA—guide RNA, Cas—CRISPR-associated protein.
Biomolecules 13 00378 g002
Figure 3. Mitochondrial proteins and complexes for which D. melanogaster models are currently available.D. melanogaster models of mitochondrial diseases include defects in OXPHOS subunits and assembly factors (listed in Table 1), defects in mtDNA maintenance (i.e., mtDNA replication machinery, nucleoid structure, and dNTP synthesis (POLG, TWINKLE, SUCLG1, MPV17, and ATAD3A)), disorders affecting mitochondrial gene expression (mitoribosomal proteins MRPS12, MRPL15, and MRPL40), and disorders affecting mitochondrial dynamics and architecture (OPA1, DRP1, MFN2, SLC25A46, and MICOS components MIC60, MIC26/27, and MIC13). Proteins are depicted using mammalian protein structures retrieved from the following PDB IDs: complex I (5LC5), complex II (1ZOY), complex III (1BGY), complex IV (2OCC), complex V (7AJD), mitoribosome (7A5F), POLG (3IKM), Twinkle (7T8C), SUCLG1 (1EUC), s-OPA1 (6QL4), MFN2 (6JFK), and DRP1 (3ZVR). No structural data are available for the proteins depicted by generic shapes in grey.
Figure 3. Mitochondrial proteins and complexes for which D. melanogaster models are currently available.D. melanogaster models of mitochondrial diseases include defects in OXPHOS subunits and assembly factors (listed in Table 1), defects in mtDNA maintenance (i.e., mtDNA replication machinery, nucleoid structure, and dNTP synthesis (POLG, TWINKLE, SUCLG1, MPV17, and ATAD3A)), disorders affecting mitochondrial gene expression (mitoribosomal proteins MRPS12, MRPL15, and MRPL40), and disorders affecting mitochondrial dynamics and architecture (OPA1, DRP1, MFN2, SLC25A46, and MICOS components MIC60, MIC26/27, and MIC13). Proteins are depicted using mammalian protein structures retrieved from the following PDB IDs: complex I (5LC5), complex II (1ZOY), complex III (1BGY), complex IV (2OCC), complex V (7AJD), mitoribosome (7A5F), POLG (3IKM), Twinkle (7T8C), SUCLG1 (1EUC), s-OPA1 (6QL4), MFN2 (6JFK), and DRP1 (3ZVR). No structural data are available for the proteins depicted by generic shapes in grey.
Biomolecules 13 00378 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Brischigliaro, M.; Fernandez-Vizarra, E.; Viscomi, C. Mitochondrial Neurodegeneration: Lessons from Drosophila melanogaster Models. Biomolecules 2023, 13, 378. https://doi.org/10.3390/biom13020378

AMA Style

Brischigliaro M, Fernandez-Vizarra E, Viscomi C. Mitochondrial Neurodegeneration: Lessons from Drosophila melanogaster Models. Biomolecules. 2023; 13(2):378. https://doi.org/10.3390/biom13020378

Chicago/Turabian Style

Brischigliaro, Michele, Erika Fernandez-Vizarra, and Carlo Viscomi. 2023. "Mitochondrial Neurodegeneration: Lessons from Drosophila melanogaster Models" Biomolecules 13, no. 2: 378. https://doi.org/10.3390/biom13020378

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