ArticlePDF Available

Metabolic and neurobehavioral disturbances induced by purine recycling deficiency in Drosophila

Authors:

Abstract and Figures

Adenine phosphoribosyltransferase (APRT) and hypoxanthine-guanine phosphoribosyltransferase (HGPRT) are two structurally related enzymes involved in purine recycling in humans. Inherited mutations that suppress HGPRT activity are associated with Lesch–Nyhan disease (LND), a rare X-linked metabolic and neurological disorder in children, characterized by hyperuricemia, dystonia, and compulsive self-injury. To date, no treatment is available for these neurological defects and no animal model recapitulates all symptoms of LND patients. Here, we studied LND-related mechanisms in the fruit fly. By combining enzymatic assays and phylogenetic analysis, we confirm that no HGPRT activity is expressed in Drosophila melanogaster , making the APRT homolog (Aprt) the only purine-recycling enzyme in this organism. Whereas APRT deficiency does not trigger neurological defects in humans, we observed that Drosophila Aprt mutants show both metabolic and neurobehavioral disturbances, including increased uric acid levels, locomotor impairments, sleep alterations, seizure-like behavior, reduced lifespan, and reduction of adenosine signaling and content. Locomotor defects could be rescued by Aprt re-expression in neurons and reproduced by knocking down Aprt selectively in the protocerebral anterior medial (PAM) dopaminergic neurons, the mushroom bodies, or glia subsets. Ingestion of allopurinol rescued uric acid levels in Aprt -deficient mutants but not neurological defects, as is the case in LND patients, while feeding adenosine or N ⁶ -methyladenosine (m ⁶ A) during development fully rescued the epileptic behavior. Intriguingly, pan-neuronal expression of an LND-associated mutant form of human HGPRT (I42T), but not the wild-type enzyme, resulted in early locomotor defects and seizure in flies, similar to Aprt deficiency. Overall, our results suggest that Drosophila could be used in different ways to better understand LND and seek a cure for this dramatic disease.
Content may be subject to copyright.
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 1 of 34
Metabolic and neurobehavioral
disturbances induced by purine recycling
deficiency inDrosophila
Céline Petitgas1,2, Laurent Seugnet3, Amina Dulac1, Giorgio Matassi4,5,
Ali Mteyrek1, Rebecca Fima1, Marion Strehaiano1, Joana Dagorret1,
Baya Chérif- Zahar1, Sandrine Marie6, Irène Ceballos- Picot2, Serge Birman1*
1Genes Circuits Rhythms and Neuropathology, Brain Plasticity Unit, CNRS, ESPCI
Paris, PSL Research University, Paris, France; 2Metabolomic and Proteomic
Biochemistry Laboratory, Necker- Enfants Malades Hospital and Paris Cité University,
Paris, France; 3Integrated Physiology of the Brain Arousal Systems (WAKING), Lyon
Neuroscience Research Centre, INSERM/CNRS/UCBL1, Bron, France; 4Dipartimento
di Scienze Agroalimentari, Ambientali e Animali, University of Udine, Udine, Italy;
5UMR “Ecology and Dynamics of Anthropogenic Systems” (EDYSAN), CNRS,
Université de Picardie Jules Verne, Amiens, France; 6Laboratory of Metabolic
Diseases, Cliniques Universitaires Saint- Luc, Université catholique de Louvain,
Brussels, Belgium
Abstract Adenine phosphoribosyltransferase (APRT) and hypoxanthine- guanine phosphoribos-
yltransferase (HGPRT) are two structurally related enzymes involved in purine recycling in humans.
Inherited mutations that suppress HGPRT activity are associated with Lesch–Nyhan disease (LND),
a rare X- linked metabolic and neurological disorder in children, characterized by hyperuricemia,
dystonia, and compulsive self- injury. To date, no treatment is available for these neurological defects
and no animal model recapitulates all symptoms of LND patients. Here, we studied LND- related
mechanisms in the fruit fly. By combining enzymatic assays and phylogenetic analysis, we confirm
that no HGPRT activity is expressed in Drosophila melanogaster, making the APRT homolog (Aprt)
the only purine- recycling enzyme in this organism. Whereas APRT deficiency does not trigger
neurological defects in humans, we observed that Drosophila Aprt mutants show both metabolic
and neurobehavioral disturbances, including increased uric acid levels, locomotor impairments,
sleep alterations, seizure- like behavior, reduced lifespan, and reduction of adenosine signaling
and content. Locomotor defects could be rescued by Aprt re- expression in neurons and repro-
duced by knocking down Aprt selectively in the protocerebral anterior medial (PAM) dopaminergic
neurons, the mushroom bodies, or glia subsets. Ingestion of allopurinol rescued uric acid levels in
Aprt- deficient mutants but not neurological defects, as is the case in LND patients, while feeding
adenosine or N6- methyladenosine (m6A) during development fully rescued the epileptic behavior.
Intriguingly, pan- neuronal expression of an LND- associated mutant form of human HGPRT (I42T),
but not the wild- type enzyme, resulted in early locomotor defects and seizure in flies, similar to Aprt
deficiency. Overall, our results suggest that Drosophila could be used in different ways to better
understand LND and seek a cure for this dramatic disease.
eLife assessment
The article looks at how dysregulated purine metabolism in mutants for the Aprt gene impacts
survival, motor, and sleep behavior in the fruit fly. Interestingly, although several deficits arise from
RESEARCH ARTICLE
*For correspondence:
serge.birman@espci.fr
Competing interest: The authors
declare that no competing
interests exist.
Funding: See page 27
Sent for Review
03 June 2023
Preprint posted
23 June 2023
Reviewed preprint posted
31 July 2023
Reviewed preprint revised
21 February 2024
Version of Record published
03 May 2024
Reviewing Editor: Gaiti Hasan,
National Centre for Biological
Sciences, India
Copyright Petitgas etal. This
article is distributed under the
terms of the Creative Commons
Attribution License, which
permits unrestricted use and
redistribution provided that the
original author and source are
credited.
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 2 of 34
dopaminergic neurons, dopamine levels are increased in Aprt mutants. Instead, the biochemical
change responsible for Aprt mutant neurobehavioral phenotypes appears to be a reduction in levels
of adenosine. This valuable study suggests that Drosophila Aprt mutants may serve as a model for
understanding Lesch–Nyhan disease (LND), caused by mutations in the human HPRT1 gene, and
may also potentially serve as a model to screen for drugs for the neurobehavioral deficits observed
in LND. The strength of evidence is solid.
Introduction
The purine salvage pathway is an essential component of cellular metabolism that allows the recovery
of free purine bases derived from the diet or from the degradation of nucleic acids and nucleotides,
thus avoiding the energy cost of de novo purine biosynthesis (Nyhan, 2014). Energy- intensive tissues,
such as cardiac muscle and brain cells, extensively use this pathway to maintain their purine levels
(Ipata, 2011; Johnson etal., 2019). The two main recycling enzymes involved in the salvage pathway
in mammals are hypoxanthine- guanine phosphoribosyltransferase (HGPRT), which converts hypoxan-
thine and guanine into IMP and GMP, respectively, and adenine phosphoribosyltransferase (APRT),
which converts adenine into AMP.
APRT and HGPRT deficiencies induce very different disorders in humans. Loss of APRT seems
to have only metabolic consequences, leading to the formation of 2,8- dihydroxyadenine crystals in
kidney, which can be fatal but is readily prevented by allopurinol treatment (Bollée et al., 2012;
Harambat et al., 2012). In contrast, highly inactivating mutations in HGPRT trigger Lesch–Nyhan
disease (LND), a rare neurometabolic X- linked recessive disorder with dramatic consequences for
child neurodevelopment (Lesch and Nyhan, 1964; Seegmiller et al., 1967). The metabolic conse-
quence of HGPRT deficiency is an overproduction of uric acid in the blood (hyperuricemia) that can
lead to gout and tophi, or nephrolithiasis (Sass etal., 1965; Kelley etal., 1967). Affected children
also develop severe neurological impairments, such as dystonia, choreoathetosis, spasticity, and a
dramatic compulsive self- injurious behavior (Nyhan, 1997; Jinnah etal., 2006; Torres etal., 2007a;
Schretlen etal., 2005; Madeo etal., 2019). They have a developmental delay from 3 to 6months
after birth, and most of them never walk or even sit without support. Xanthine oxidase inhibitors, such
as febuxostat or allopurinol, are given to patients after diagnosis to decrease their uric acid levels and
prevent the formation of urate crystals in kidney, which can lead to renal failure (Kelley etal., 1967;
Torres etal., 2007a; Lahaye et al., 2014). However, there is as yet no treatment to alleviate the
neurological symptoms of LND (Torres and Puig, 2007b; Jinnah etal., 2010; Madeo etal., 2019).
To date, the causes of neurobehavioral troubles in LND are still not elucidated (Jinnah etal., 2010;
Bell etal., 2016). The most favored hypothesis is a dysfunction of the brain’s basal ganglia, and partic-
ularly of its dopaminergic pathways (Baumeister and Frye, 1985; Visser etal., 2000; Nyhan, 2000;
Saito and Takashima, 2000; Egami etal., 2007). Indeed, analyses revealed a marked loss of dopa-
mine (DA) (Lloyd etal., 1981; Ernst etal., 1996) and DA transporters (Wong etal., 1996) in the basal
ganglia of LND patients. DA deficits have also been reported in HGPRT knockout rodents, but without
motor or behavioral defects (Finger etal., 1988; Dunnett etal., 1989; Jinnah etal., 1993; Jinnah
etal., 1994; Meek etal., 2016). Recent studies reported that HGPRT deficiency disrupts prolifer-
ation and migration of developing midbrain DA neurons in mouse embryos, arguing for a neurode-
velopmental syndrome (Witteveen etal., 2022). This could result from ATP depletion and impaired
energy metabolism (Bell etal., 2021) or an overactivation of de novo purine synthesis, leading to the
accumulation of potentially toxic intermediates of this pathway (López, 2008; López et al., 2020).
Pharmacological models have also been developed by injecting the neurotoxin 6- hydroxydopamine
into neonatally rat brains, which induced a self- mutilation behavior in response to DA agonist admin-
istration in adulthood. However, these models are of limited utility as they do not reproduce the basic
genetic impairment of LND (Breese etal., 1990; Knapp and Breese, 2016; Bell etal., 2016).
New animal models are therefore needed to study LND pathogenesis and find efficient therapeutic
molecules. The fruit fly Drosophila melanogaster presents many advantages for translational studies
and drug discovery (Fernández- Hernández etal., 2016; Perrimon etal., 2016; Papanikolopoulou
etal., 2019). Although the importance of this invertebrate model for studying rare human genetic
diseases is now recognized (Oriel and Lasko, 2018), a Drosophila LND model has not yet been devel-
oped to our knowledge. This is probably due to the fact that no HGPRT activity has been detected in
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 3 of 34
this organism (Miller and Collins, 1973; Becker,
1974a; Becker, 1974b). However, an ortholog
of APRT is expressed in Drosophila (Johnson
et al., 1987), encoded by the Aprt gene. It is
therefore possible that part of the functions of
human HGPRT, and in particular those essential
for nervous system development and neurophysi-
ology, are endorsed by Aprt in Drosophila.
Here, we studied the effects of Aprt deficiency
on purine metabolism, lifespan, and various adult
fly behaviors, including spontaneous and startle-
induced locomotion, sleep, and seizure- like bang
sensitivity (BS). We show that Aprt is required
during development and in adult stage for many
aspects of Drosophila life, and that its activity is of
particular importance in subpopulations of brain
dopaminergic neurons and glial cells. Lack of
Aprt appears to decrease adenosinergic signaling
and induces both metabolic and neurobehavioral
symptoms in flies, as is the case with HGPRT in
humans. We also find that expression of an LND- associated mutant form of HGPRT, but not the wild-
type enzyme, in Drosophila neurons, induced neurobehavioral impairments similar to those of Aprt-
deficient flies. Such a potential toxic gain- of- function effect of mutated HGPRT had not yet been
demonstrated in an animal model .
Results
Evolution of HGPRT proteins
The pathways of purine anabolism/catabolism and recycling have been closely conserved between
Drosophila melanogaster and humans (Figure1—figure supplement 1): all genes related to purine
metabolism have homologs in both species, except for the human HPRT1 gene encoding HGPRT
(step 13 in Figure1—figure supplement 1), which has no ortholog in flies, and the lack of urate
oxidase in humans (step 20). In accordance with pioneering reports from about 50years ago (Miller
and Collins, 1973; Becker, 1974a; Becker, 1974b), we confirmed that no HGPRT enzymatic activity
can be detected in extracts of wild- type D. melanogaster (see below Table 2), using either hypoxan-
thine or guanine as substrate in the reaction medium. This intriguing observation prompted us to carry
out a more precise analysis of the evolution of HGPRT.
HGPRT proteins are ancient, for they are present in both bacteria and archaea. However, the analysis
of the phyletic distribution of HGPRT proteins revealed their striking rareness in insecta. This conclu-
sion is based on PSI- Blast sequence similarity searches on the NCBI Insecta database (taxid: 6960,
50557). Phylogenetic analysis showed that the only 11 HGPRT proteins found in Insecta cluster mainly
with bacteria, but also with fungi, apicomplexa, and acari (Figure1—figure supplement 2, red font,
see legend for details). These and further evidence support the acquisition of HGPRT in a few insecta
species by horizontal gene transfer (G. Matassi, unpublished observations). In particular, HGPRT has
no homolog in Drosophilidae, with the potential exception of a single species, Drosophila immigrans,
in which our most recent PSI- BLAST analysis detected one hit (accession KAH8256851.1, annotated
as hypothetical protein). Yet this sequence is 100% identical to the HGPRT of the Gammaproteobac-
terium Serratia marcescens. A phylogenetic analysis showed that the D. immigrans HGPRT clusters
with the Serratia genus (Figure1—figure supplement 3), suggesting either a contamination of the
sequenced sample or a very recent horizontal gene transfer event. The second scenario is more likely
since the corresponding nucleotide sequences differ by five synonymous substitutions (out of 534
positions). We also carried out structural similarity searches against the RCSB Protein Data Bank repos-
itory using the human HGPRT structure as query (PDB identifiers: 5HIA or 1Z7G). This analysis did not
identify any protein with a divergent sequence and relevant similarity with HGPRT 3D structure in D.
melanogaster, consistent with the lack of HGPRT enzymatic activity in this organism.
Table 1. Aprt activity in wild- type and Aprt-
deficient flies.
Genotypes Sex
Aprt activity
(nmol/min/
mg prot)
Wild type Males 1.32 ± 0.17
Females 2.77 ± 0.27
Aprt5/Aprt5Males and females 0.04 ± 0.02
Aprt5/Df(3L)ED4284 Males 0.02 ± 0.01
da/+ Males 2.78 ± 0.41
da>AprtRNAi Males 0.10 ± 0.01
AprtRNAi/+ Males 2.16 ± 0.37
The online version of this article includes the following
source data for table 1:
Source data 1. Source data for Table1.
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 4 of 34
Drosophila lacking Aprt activity have a shortened lifespan
Both the phylogeny and enzymatic assays therefore strongly suggest that Aprt (Figure1—figure
supplement 1, step 12) is the only recycling enzyme of the purine salvage pathway in Drosophila.
To assess the importance of purine recycling in brain development and function in this organism, we
analyzed the phenotypes induced by a deficiency in Aprt. The Aprt5 mutant was originally gener-
ated by chemical mutagenesis followed by selection of flies resistant to purine toxicity (Johnson and
Friedman, 1981; Johnson and Friedman, 1983). Enzymatic assays confirmed a strong reduction in
Aprt activity in extracts of heterozygous Aprt5/+ mutantsand its absence in homozygous and hemizy-
gous (Aprt5/Df(3L)ED4284) mutants (Figure1—figure supplement 4 and Table1). Sequencing of the
Aprt5 cDNA (Figure1—figure supplement 5A) indicated that the Aprt5 mRNA codes for a protein
with several amino acid changes compared to D. melanogaster wild- type Aprt, modifying in partic-
ular three amino acid residues that have been conserved in the Aprt sequences from Drosophila to
humans (Figure1—figure supplement 5B). These mutations are likely to be responsible for the loss
of enzymatic activity. The homozygous Aprt5 mutants are considered viable because they develop and
reproduce normally. However, we observed that these mutants have a significantly reduced longevity,
their median lifespan being only 38d against 50d for wild- type flies (p<0.001) (Figure1A).
Uric acid levels are increased in Aprt5 mutants and rescued by
allopurinol
In humans, one of the consequences of HGPRT deficiency on metabolism is the overproduction of uric
acid (Harkness etal., 1988; Fu etal., 2015). We assayed the levels of purine metabolites by HPLC
and found that the level of uric acid was substantially increased by 37.7% on average in Drosophila
Aprt5 mutant heads (p<0.01) (Figure1B). We then tried to rescue uric acid content in flies by providing
allopurinol in the diet, as is usually done for LND patients. Allopurinol is a hypoxanthine analog and
a competitive inhibitor of xanthine oxidase, an enzyme that catalyzes the oxidation of xanthine into
uric acid (Figure1—figure supplement 1, step 19). Remarkably, the administration of 100μg/ml allo-
purinol during 5d decreased uric acid levels to a normal concentration range in Aprt5 mutant heads
(Figure1B). Therefore, quite similarly to HGPRT deficit in humans, the lack of Aprt activity in flies
increases uric acid levels and this metabolic disturbance can be prevented by allopurinol.
Aprt deficiency decreases motricity in young flies
LND patients present dramatic motor disorders that prevent them for walking at an early age. To
examine whether a deficiency in the purine salvage pathway can induce motor disturbance in young
flies, we monitored the performance of Aprt- null mutants in startle- induced negative geotaxis (SING),
a widely used paradigm to assess climbing performance and locomotor reactivity (Feany and Bender,
2000; Friggi- Grelin et al., 2003; Riemensperger et al., 2013; Sun et al., 2018). Strikingly, Aprt5
mutant flies showed a very early SING defect starting from 1day after eclosion (d a.E.) (performance
index [PI] = 0.73 vs 0.98 for wild- type flies, p<0.001) that was more pronounced at 8 d a.E. (PI = 0.51
vs 0.96 for the wild- type flies, p<0.001). The fly locomotor performance did not further decline after-
wards until 30 d a.E. (Figure1C). Df(3L)ED4284 and Df(3L)BSC365 are two partially overlapping small
genomic deficiencies that uncovers Aprt and several neighbor genes. Hemizygous Aprt5/Df(3L)ED4284
or Aprt5/Df(3L)BSC365 flies also displayed SING defects at 10 d a.E. (PI = 0.71and 0.68, respectively,
compared to 0.97 for wild- type flies, p<0.01) (Figure 1—figure supplement 6). In contrast to its
beneficial effect on uric acid levels, we observed that allopurinol treatment did not improve the loco-
motor ability of Aprt5 mutant flies, either administered 5d before the test or throughout the devel-
opment (Figure1D and E). This is comparable to the lack of effect of this drug against neurological
defects in LND patients.
To confirm the effect of Aprt deficiency on the SING behavior, we used an UAS- AprtRNAi line that
reduced Aprt activity by more than 95% when expressed in all cells with the da- Gal4 driver (Table1).
These Aprt knock- down flies also showed a strong SING defect at 10 d a.E. (PI = 0.62 against 0.97
and 0.85 for the driver and UAS- AprtRNAi alone controls, respectively, p<0.001), like the Aprt5 mutant
(Figure1F). Next, we used tub- Gal80ts, which inhibits Gal4 activity at permissive temperature (McGuire
etal., 2003), to prevent Aprt knockdown before the adult stage. Tub- Gal80ts; da- Gal4>AprtRNAi flies
raised at permissive temperature (18°C) did not show any locomotor impairment (Figure1G, white
bars). However, after being transferred for 3d (between 7 and 10 d a.E.) at a restrictive temperature
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 5 of 34
Figure 1. Aprt deciency shortens lifespan and induces metabolic and neurobehavioral defects. (A) Aprt5 mutant ies have a reduced lifespan
compared to wild- type ies (median lifespan: 38 and 50d, respectively). Three independent experiments were performed on 150males per genotype
with similar results and a representative experiment is shown. Log- rank test (***p<0.001). (B) HPLCproles on head extracts revealed an increase in
uric acid levels in Aprt5 mutant ies. Administration of 100μg/ml allopurinol for 5d before the test rescued uric acid levels. Mean of three independent
Figure 1 continued on next page
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 6 of 34
(30°C) that inactivates Gal80, the same flies demonstrated a similar SING defect as Aprt5 mutants
(PI = 0.63 compared to 0.94 and 0.83 for the driver and RNAi alone controls, respectively, p<0.001)
(Figure1G, yellow bars with dots). This demonstrates that Aprt inactivation at the adult stage only is
sufficient to alter SING performance in Drosophila, strongly suggesting that this genotype does not
result from developmental defects.
Cell specificity of Aprt requirement for startle-induced locomotion
We then tried to identify the neural cells in which Aprt is required to ensure a normal locomotor reac-
tivity in young flies by expressing AprtRNAi with selective Gal4 drivers. Expression in all neurons with
elav- Gal4 led to decreased locomotor performance in the SING test (PI = 0.68 at 10 d a.E., vs 0.90
and 0.86 for the driver and RNAi controls, respectively, p<0.05) (Figure2A), which was comparable
to the effects observed with the Aprt5 mutant or after ubiquitous expression of the RNAi. To confirm
the role of neuronal Aprt in locomotor control, we generated a UAS- Aprt line, which allowed for a
substantial increase in Aprt expression and activity (Figure2—figure supplement 1). We then found
that re- expressing Drosophila Aprt selectively in neurons in Aprt5 background partially rescued the
SING phenotype of the null mutant (PI = 0.70 vs 0.52 for driver and UAS- Aprt controls in Aprt5 back-
ground, p<0.05) (Figure2B).
Furthermore, Aprt knockdown in all glial cells with repo- Gal4, or in sub- populations of glial cells
that express the glutamate transporter Eaat1 with Eaat1- Gal4, which includes astrocyte- like glia,
cortex glia, and some subperineurial glia (Rival et al., 2004; Mazaud etal., 2019), also led to a
lower SING performance of 10- day- old flies (PI = 0.72 for repo- Gal4 vs 0.91 for both controls, p<0.05,
and PI = 0.56 for Eaat1- Gal4 vs 0.77, p<0.05, and 0.88, p<0.01, for the driver and RNAi controls,
respectively) (Figure2C and D). In contrast, MZ0709- Gal4 (Doherty etal., 2009) and NP6520- Gal4
(Awasaki et al., 2008) that selectively target the ensheathing glia did not induce any significant
locomotor defects when used to express the Aprt RNAi (Figure2—figure supplement 2). Notice-
ably, re- expressing Aprt with Eaat1- Gal4 in the Aprt5 background did not rescue the SING pheno-
type (Figure2E), in contrast to the positive effect of neuronal re- expression (Figure2B). Overall this
experiments performed on 40 ies per genotype. One- way ANOVA with Tukey’s post hoc test for multiple comparisons (*p<0.05; **p<0.01; ns: not
signicant). (C–E)Effect on climbing ability. (C)Aprt5 mutants shows early defects in the startle- induced negative geotaxis (SING) paradigm that monitors
locomotor reactivity and climbing performance. This decit was already obvious at 1 day after eclosion (d a.E.) and further decreased up to 8 d a.E., after
which it did not change signicantly up to 30 d a.E. Mean of three independent experiments performed on 50 ies per genotype. Unpaired Student’s
t- test (**p<0.01; ***p<0.001). (D, E)Administration of allopurinol does not rescue the motricity defects of Aprt- decient mutants. Feeding the Aprt5
mutants with allopurinol (100μg/ml) either in adults 5d before the test (D)or throughout all developmental stages (E)did not alter the defects observed
in SING behavior. Results of one experiment performed on 50 ies per genotype at 10 d a. E. Unpaired Student’s t- test (***p<0.001). (F)Downregulating
Aprt by RNAi in all cells (da>AprtRNAi) also led to an early impairment in climbing responses in the SING assay at 10 d a.E. compared to the driver (da/+)
and effector (AprtRNAi/+) only controls. Mean of three independent experiments performed on 50 ies per genotype. One- way ANOVA with Tukey’s
post hoc test for multiple comparisons (***p<0.001). (G)Adult- specic inactivation of Aprt (tub- Gal80ts; da- Gal4>AprtRNAi) decreased startle- induced
climbing abilities in the SING paradigm, suggesting that the locomotor impairment induced by Aprt deciency is not caused by a developmental effect.
Flies were raised at permissive temperature (18°C) in which Gal80ts suppressed Gal4- controlled AprtRNAi expression and were shifted from 18 to 30°C
for 3d before the test (between 7 and 10 d a.E.) to activate transgene expression. Mean of three independent experiments performed on 50 ies per
genotype. Two- way ANOVA with Sidak’s post hoc test for multiple comparisons (***p<0.001; ns: not signicant).
The online version of this article includes the following source data and gure supplement(s) for gure 1:
Source data 1. Source data for Figure1A–G.
Figure supplement 1. Comparison of purine metabolism pathways in Drosophila and humans.
Figure supplement 2. Urooted maximum likelihood phylogeny of hypoxanthine- guanine phosphoribosyltransferase (HGPRT) proteins (189 taxa, 130
sites).
Figure supplement 3. Urooted maximum likelihood phylogeny of HPRT proteins (20 taxa, 177 sites).
Figure supplement 4. Lack of Aprt enzymatic activity in the Aprt5 mutant.
Figure supplement 4—source data 1. Source data for Figure1—figure supplement 4.
Figure supplement 5. Alignment of wild- type and mutant Aprt cDNAs and predicted protein sequences.
Figure supplement 6. Startle- induced negative geotaxis (SING) behavior of hemizygous Aprt mutant ies.
Figure supplement 6—source data 1. Source data for Figure1—figure supplement 6.
Figure 1 continued
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 7 of 34
Figure 2. Aprt knockdown in neurons or glial cells disrupts startle- induced locomotion in Drosophila. (A)AprtRNAi expression in all neurons with elav-
Gal4 decreased startle- induced negative geotaxis (SING) performance in 10- day- old ies. (B)Pan- neuronal expression of Drosophila Aprt with the
UAS- Gal4 system partially rescued the locomotor response of Aprt5 mutants. (C, D)Downregulation of Aprt expression in all glia with repo- Gal4 (C)or
in glial cell that express the glutamate transporter Eaat1 with Eaat1- Gal4 (D)also altered SING performances. (E)Aprt re- expression in glial cells with
Eaat1- Gal4 driver did not rescue the climbing response of Aprt5 mutants. Results of three or four independent experiments performed on 50 ies per
genotype at 10 days after eclosion (d a.E.). One- way ANOVA with Tukey’s post hoc test for multiple comparisons (*p<0.05; **p<0.01; ns: not signicant).
The online version of this article includes the following source data and gure supplement(s) for gure 2:
Source data 1. Source data for Figure2A–E.
Figure supplement 1. Transgenic expression of Drosophila Aprt.
Figure supplement 1—source data 1. Source data for Figure2—figure supplement 1.
Figure supplement 2. Downregulation of Aprt expression in the ensheathing glia does not alter locomotor performances.
Figure supplement 2—source data 1. Source data for Figure2—figure supplement 2.
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 8 of 34
suggests that Aprt is required both in neurons and glia subsets to ensure a normal SING performance
in young flies, and that neuronal but not glial Aprt re- expression is sufficient to restore a partially
functional locomotor behavior.
To identify the neuronal subpopulations in which Aprt is required to ensure proper locomotor
responses in young flies, we first tested dopaminergic drivers since we previously showed that DA
neurons play an important role in the control of the SING behavior (Riemensperger etal., 2011;
Riemensperger et al., 2013; Sun etal., 2018). Aprt knockdown targeted to these neurons with
TH- Gal4 did not induce significant impairments (Figure 3A). This driver expresses in all brain DA
neurons except a large part of the protocerebral anterior medial (PAM) cluster (Friggi- Grelin etal.,
2003; Mao and Davis, 2009; Pech etal., 2013). In contrast, downregulating Aprt with the TH- Gal4,
R58E02- Gal4 double driver, which labels all DA neurons including the PAM cluster (Sun etal., 2018),
induced a quite similar locomotor defect as did pan- neuronal Aprt knockdown (PI = 0.72 vs 0.96 and
0.93 for the driver and RNAi controls, respectively, p<0.001) (Figure3B). Besides, downregulating
Aprt in a majority of the serotonergic neurons with TRH- Gal4 (Cassar etal., 2015) did not induce a
SING defect (Figure3C).
These results suggest that some DA neurons in the PAM cluster are specifically involved in the
locomotor impairments induced by Aprt deficiency. It has been previously shown in our laboratory
that inhibiting DA synthesis in a subset of 15 PAM DA neurons cluster was able to alter markedly
SING performance in aging flies (Riemensperger etal., 2013). We and others also reported that
the degeneration or loss of PAM DA neurons is involved in the SING defects observed in several
Drosophila models of Parkinson’s disease (Riemensperger etal., 2013; Bou Dib etal., 2014; Ta s
etal., 2018; Pütz et al., 2021). Here we found that expressing AprtRNAi in the PAM cluster with
R58E02- Gal4 reproduced the same motor disturbance as pan- neuronal expression (PI = 0.74 vs
0.96, p<0.001, and 0.85, p<0.05, for the driver and RNAi controls, respectively) (Figure3D), and
this result was confirmed by using two other PAM drivers (NP6510- Gal4 – expressing only in 15
neurons – and R76F05- Gal4) (Figure3E and F). This strongly suggests that purine recycling defi-
ciency compromises the correct functioning of these neurons, leading to a defective startle- induced
locomotion.
Because PAM DA neurons innervate the horizontal lobes of the mushroom bodies (Liu etal., 2012;
Riemensperger etal., 2013), and because this structure has been shown to be involved in the control
of startle- induced climbing (Riemensperger etal., 2013; Bou Dib etal., 2014; Sun et al., 2018),
we also inactivated Aprt in mushroom body neurons with 238Y- Gal4 that strongly expresses in that
structure (Aso etal., 2009). Interestingly, this driver did induce a locomotor reactivity impairment
(PI = 0.70 vs 0.89 for both controls, p<0.01) (Figure 3G), and the same result was observed with
VT30559- Gal4, which is a very specific driver for all the mushroom body intrinsic neurons (Plaçais
etal., 2017; Figure3H). Overall, these results show that normal expression of the SING behavior
depends on Aprt expression in the PAM and mushroom body neurons in Drosophila.
Sleep disturbances induced by Aprt deficiency
Both the mushroom body and subpopulations of PAM DA neurons are known to be regulators of
sleep in Drosophila (Nall etal., 2016; Artiushin and Sehgal, 2017). The fact that Aprt deficiency in
some of these cells impaired locomotor regulation prompted us to monitor the spontaneous loco-
motion and the sleep pattern of Aprt mutants. Compared to controls, Aprt- deficient flies did not
have an altered circadian activity profile (Figure4—figure supplement 1), nor any difference in total
spontaneous locomotor activity during the day, as quantified by the number of infrared beam cuts
(events) per 30min, but they showed a 26.2% increase in total activity during the night (p<0.001)
(Figure4A). As usual, a sleep bout was defined as 5min or more of fly immobility (Huber etal., 2004),
and we checked that wild- type and Aprt mutant flies that did not move for 5min were indeed asleep
because they were less sensitive to mild mechanical stimulation (Figure4—figure supplement 2).
We found that Aprt5 mutants slept significantly less than wild- type flies and that it was the case both
during day and night (Figure4B and C). These mutants indeed showed a reduced walking speed
(Figure4D) and a smaller average sleep bout duration (Figure4E), indicating a difficulty to maintain
sleep. The reduced speed does not seem to be caused by a decreased energetic metabolism since
we could not detect different ATP levels in head and thorax of Aprt5 mutants compared to wild- type
flies (Figure4—figure supplement 3).
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 9 of 34
Figure 3. Aprt downregulation in dopamine (DA) neurons of the protocerebral anterior medial (PAM) cluster and
in mushroom body neurons impairs startle- induced locomotion. (A)RNAi- mediated Aprt inactivation in brain
DA neurons except a large part of the PAM cluster with the TH- Gal4 driver did not lead to locomotor defects in
the startle- induced negative geotaxis (SING) assay. (B)In contrast, Aprt knockdown in all dopaminergic neurons
Figure 3 continued on next page
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 10 of 34
Interestingly, RNAi- mediated downregulation of Aprt selectively in neurons (elav>AprtRNAi flies)
also significantly decreased sleep duration during day and night (Figure 4F), whereas glial- only
expression (repo>AprtRNAi) had no effect (Figure4—figure supplement 4A). Expressing the RNAi in
both neurons and glia (elav; repo>AprtRNAi) had similar effects as in neurons alone (Figure4—figure
supplement 4B). Quantification of total sleep in these experiments, and the total amount of day
and night sleep, are shown in Figure4G and Figure4—figure supplement 4C and D, respectively.
Overall, these results suggest that Aprt expression is selectively needed in neurons for a normal sleep
pattern in Drosophila.
Brain DA synthesis and levels are increased in Aprt-deficient flies
Since we found that induced locomotion and sleep, two behaviors controlled by DA neurons in
Drosophila, were altered in Aprt- deficient flies, we expected brain DA levels to be reduced in these
mutants, as is the case in the basal ganglia of LND patients. We therefore carried out comparative
immunostaining for tyrosine hydroxylase (TH), the specific enzyme for DA synthesis (Friggi- Grelin
etal., 2003; Riemensperger etal., 2011), on dissected adult brains from wild- type and Aprt5 mutant
flies. However, the global TH protein level appeared not to be decreased, but relatively increased by
17.5% in the Aprt5 mutant brain (p<0.01), in particular around the mushroom bodies, a structure that
receives dense dopaminergic projections (Figure5A and B). Moreover, DA immunostaining carried
out on whole- mount dissected brains revealed a similarly increased level of this neuromodulator in
Aprt5 flies, by 17.0% on average in the entire brain (p<0.01), but not specifically in the mushroom
bodies or another part of the brain (Figure5C and D). We also found that the transcript levels of
DTH1, encoding the TH neuronal isoform in Drosophila, were increased in Aprt5 mutants compared
to wild- type flies (Figure5E), and, conversely, decreased when Aprt was overexpressed ubiquitously
(Figure 5F). Western blot experiments further indicated that DTH1 protein levels are increased in
Aprt5 compared to controls (Figure5G and H). This indicates that disruption of the purine salvage
pathway does not impede DA synthesis and levels in the Drosophila brain, which are instead slightly
increased. We therefore searched for another neurotransmitter system that could be affected by Aprt
deficiency.
Interactions between Aprt and the adenosinergic system
Aprt catalyzes the conversion of adenine into AMP, and AMP breakdown by the enzyme 5-nucle-
otidase produces adenosine, primarily in the extracellular space. Adenosine then acts as a wide-
spread neuromodulator in the nervous system by binding to adenosine receptors. We therefore
suspected that adenosine level could be reduced in the absence of Aprt activity, leading to alter-
ations in adenosinergic neurotransmission. Indeed, we found a significant decrease in adenosine level
either in whole flies (by 61.0% on average, p<0.01) or in heads (by 48.0%, p<0.05) in Aprt5 mutants
(Figure6A). We then examined the consequences of this reduction on molecular components of the
adenosinergic system, namely G protein- coupled adenosine receptors and equilibrative nucleoside
transporters (ENTs), which carry out nucleobase and nucleoside uptake of into cells. Only one seven-
transmembrane- domain adenosine receptor, AdoR, is present in Drosophila, which is very similar to
mammalian adenosine receptors (Dolezelova etal., 2007; Brody and Cravchik, 2000), and three
putative equilibrative nucleoside transporters (Ent1- 3) have been identified (Sankar etal., 2002) but
only one, Ent2, showed nucleoside transport activity (Machado etal., 2007). Interestingly, Ent2 has
including the PAM cluster with the TH- Gal4, R58E02- Gal4 double- driver led to a decrease in SING performance.
(C)Aprt downregulation in serotonergic neurons with TRH- Gal4 did not alter startle- induced climbing response of
the ies. (D–F) Aprt knockdown selectively in DA neurons of the PAM cluster using either R58E02- Gal4 (D),NP6510-
Gal4 (E),or R76F05- Gal4 (F)signicantly decreased climbing performance. (G, H)Aprt knockdown in all the
mushroom body intrinsic neurons (Kenyon cells) with 238Y- Gal4 (G)or VT30559- Gal4 (H)also led to a decrease
in SING performance. Results of three or four independent experiments performed on 50 ies per genotype at
10 days after eclosion (d a.E.). One- way ANOVA with Tukey’s post hoc test for multiple comparisons (*p<0.05;
**p<0.01; ***p<0.001; ns: not signicant).
The online version of this article includes the following source data for gure 3:
Source data 1. Source data for Figure3A–H.
Figure 3 continued
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 11 of 34
Figure 4. Aprt- decient ies sleep less and walk slower than wild- type ies. (A)Quantication of total spontaneous locomotor activity during day and
night over ve light- dark (LD) cycles. Aprt5 mutants show no difference in spontaneous locomotion with wild- type ies during the day but a higher
activity at night. Three independent experiments were performed on 32 ies per genotype and mean ± SEM was plotted. Unpaired Student’s t- test
(***p<0.001; ns: not signicant). (B)Sleep pattern during a typical 24hr LD cycle showing that the total amount of sleep is smaller during day and
night in Aprt5 mutants compared to wild- type ies. (C)Quantication of day (ZT1- 12), night (ZT13- 24), and total sleep in Aprt5 mutants. ZT, zeitgeber.
(D)Locomotion speed during waking is reduced in Aprt5 mutants. (E)The average sleep bout duration is also decreased, indicating that Aprt5 mutants
have a difculty to maintain sleep. (F)Sleep pattern of elav>AprtRNAi ies, showing that knockdown of Aprt in all neurons led to sleep reduction during
the night, similarly to the mutant condition, and an even more profound sleep defect during the day. (G)Quantication of total amount of sleep when
Aprt was downregulated in all neurons (elav- Gal4), all glial cells (repo- Gal4), and both neurons and glial cells (elav- Gal4; repo- Gal4). Except for glia only,
Figure 4 continued on next page
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 12 of 34
been shown to be enriched in the mushroom bodies and its transcript level to be highly elevated in
AdoR mutant flies (Knight etal., 2010). In Aprt5 background, we also observed a strong increase
in Ent2 mRNAs (2.3- fold higher than wild- type flies, p<0.001), but no noticeable effect on AdoR
transcript level (Figure6B). The increased expression of Ent2 could correspond to a compensatory
response to adenosine shortage in Aprt5 mutants.
The AdoR receptor is highly expressed in adult fly heads and its ectopic overexpression leads to
early larval lethality (Dolezelova etal., 2007). In contrast, a null mutant of this receptor, AdoRKGex,
in which the entire coding sequence is deleted, is fully viable (Wu etal., 2009). This enabled us to
examine the consequences of a complete lack of AdoR on purine recycling in adult flies. We found that
Aprt transcripts were decreased by 29.5% on average (p<0.001) in AdoRKGex mutant heads (Figure6C,
left panel), while Aprt activity was even more decreased by 78.4% (p<0.001) compared to wild- type
flies (Figure 6C, right panel). This effect likely results from the much increased level of adenosine
uptake in the AdoR mutants (Knight et al., 2010), which would downregulate Aprt activity by a
feedback mechanism. Overall, these data indicate that extracellular adenosine levels must be strongly
decreased in Aprt mutant flies, both from the general reduction in adenosine levels and the increased
expression of the Ent2 transporter. Although AdoR expression is not affected, AdoR signaling is there-
fore expected to be significantly reduced in Aprt mutants.
Aprt mutants show epilepsy-like seizure behavior
A number of Drosophila mutants with disrupted metabolism or neural function show increased
susceptibility to seizure and paralysis following strong mechanical or electrophysiological stimulation
(Fergestad et al., 2006; Parker et al., 2011; Kroll et al., 2015). These mutants are called bang
sensitive (BS) and commonly used as models of epileptic seizure (Song and Tanouye, 2008). Here we
checked whether Aprt deficiency could trigger a BS phenotype. We observed that aged Aprt mutants
(at 30 d a.E.) recovered slowly after a strong mechanical stimulation applied by vortexing the vial for
10s at high speed. These flies took on average 17.3s to recover and get back on their legs compared
to 2.5s for wild- type flies of the same age (p<0.01) (Figure7A; see also Figure7—videos 1 and 2
). Some of the mutant flies appeared more deeply paralyzed as they did not spontaneously recover
unless the vial was stirred, so their recovery time could not be scored. Hemizygous flies of the same
age containing the Aprt5 mutation over two partially overlapping genomic deficiencies covering Aprt
also showed a BS behavior, with an average longer recovery time of 28.9s and 33.2s, respectively
(p<0.001) (Figure7B).
We then downregulated Aprt by RNAi in all cells with the da- Gal4 driver to check if this could also
induce BS behavior. As shown in Figure7C, da>AprtRNAi flies at 30 d a.E. indeed displayed seizure
after mechanical shock, quite similar to that of the Aprt5 mutants (22.7s recovery time on average
compared to 1.55s and 0.72 s for the driver and RNAi controls, respectively, p<0.05). In contrast,
inactivating Aprt selectively in neurons (elav- Gal4), glial cells (repo- Gal4), or muscles (24B- Gal4) did
not induce a BS phenotype (Figure7—figure supplement 1). This suggests that the BS phenotype
requires Aprt knockdown in other cells or in several of these cell types. Finally, in contrast to the SING
defect, we observed that adult- specific Aprt knockdown in all cells with da- Gal4 for 3d did not trigger
the resulting effect was a signicant sleep reduction. For sleep and locomotion speed, means ± SEM were plotted. Unpaired Student’s t- test (C–E)and
one- way ANOVA with Tukey’s post hoc test for multiple comparisons (G) (*p<0.05; **p<0.01; ***p<0.001; ns: not signicant).
The online version of this article includes the following source data and gure supplement(s) for gure 4:
Source data 1. Source data for Figure4A–G.
Figure supplement 1. Daily locomotor activity proles of wild- type and Aprt5 mutant ies.
Figure supplement 2. Sensitivity of resting ies to mild mechanical stimulation.
Figure supplement 2—source data 1. Source data for Figure4—figure supplement 2.
Figure supplement 3. ATP levels are not altered in head and thorax of Aprt5 ies compared to the wild- type.
Figure supplement 3—source data 1. Source data for Figure4—figure supplement 3.
Figure supplement 4. Sleep patterns of ies with cell- specic Aprt deciency.
Figure supplement 4—source data 1. Source data for Figure4—figure supplement 4.
Figure 4 continued
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 13 of 34
Figure 5. Aprt deciency increases dopamine (DA) synthesis and content in the Drosophila brain.
(A)Representative confocal projections of tyrosine hydroxylase (TH)- immunostained whole- mount adult brains
from wild- type ies and Aprt5 mutants. MB: mushroom body. Scale bars: 100μm. (B)Quantication of TH
immunouorescence intensity normalized to the controls in the entire brain. 4–6 brains were dissected per
Figure 5 continued on next page
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 14 of 34
a BS behavior (Figure7D), indicating that the BS requires a longer downregulation of the gene or
might be the consequence of a developmental defect.
Administration of adenosine or N6-methyladenosine to Aprt-deficient
flies prevents seizure
Drosophila disease models are advantageously tractable for drug screening in vivo (Fernández-
Hernández etal., 2016; Perrimon et al., 2016). We thus administered several compounds related
to purine metabolism to Aprt5 flies to check if they can rescue neurobehavioral impairments (loco-
motor defects and seizure). Feeding allopurinol at the same concentration used for uric acid rescue
(100μg/ml, Figure1B), either in adults 5d before the test or throughout all developmental stages,
did not alter the BS phenotype (Figure7—figure supplement 2), as was the case for the SING assay
(Figure1D and E). Similarly in humans, it has been shown that the daily intake of allopurinol, even
from infancy, does not mitigate the neurobehavioral impairments in LND patients (Marks etal., 1968;
Torres etal., 2007a; Jinnah etal., 2010; Madeo etal., 2019).
Then, we tried to supplement Aprt mutants with various purine compounds, including adenine,
hypoxanthine, adenosine, and N6- methyladenosine (m6A), at 100 or 500 μM, either in adult stage
5d before the test or throughout larval development plus 5d before the test. None of these drugs
was able to rescue the SING defect (Figure7—figure supplement 3). In contrast and interestingly,
administration of 500μM adenosine or m6A during development rescued the BS phenotype of Aprt5
mutants (Figure7E and F). This further indicates that different mechanisms underpin SING alteration
and BS behavior in Aprt mutants and provide another evidence that the BS may be caused by a devel-
opmental defect. The results also suggest that the lower adenosine levels of Aprt mutant flies could
be at the origin of their BS.
Neuronal expression of mutant HGPRT induces early locomotor defects
and seizure behavior
In order to potentially develop another Drosophila model mimicking LND conditions, we generated
new transgenic UAS lines to express in living flies either the human wild- type HGPRT (HGPRT- WT) or
a pathogenic LND- associated mutant form of this protein (HGPRT- I42T), both isoforms being inserted
at the same genomic docking site. These lines were validated by showing that they are transcribed at
similar levels (Figure8A and B). Enzymatic assays on adult extracts of flies expressing the wild- type
form HGPRT- WT in all cells with da- GAL4 showed significant HGPRT enzyme activity, while no activity
experiment and genotype, and 6 independent experiments were performed (**p<0.01). (C)Representative
confocal projections of DA immunostaining in whole- mount adult brains of wild- type and Aprt5 mutants. Scale
bars: 100μm. (D)Quantication of DA immunouorescence intensity over the entire brain showed a slight increase
in DA content in Aprt5 mutants compared to wild- type controls. Six brains were dissected per experiment and
genotype, and three independent experiments were performed. Unpaired Student’s t- test (**p<0.01). (E) mRNA
levels of TH neuronal form DTH1 are increased in Aprt5 mutant heads compared to wild- type ies. Results of six
independent RT- qPCR experiments carried out on 3–4 different RNA extractions from 20 to 30 male heads per
genotype. Unpaired Student’s t- test (**p<0.01). (F)Conversely, overexpressing Aprt in all cells with the da- Gal4
driver (da>Aprt) reduced mRNA level of DTH1 in heads compared to the driver (da/+) and effector (Aprt/+)
controls. Mean of three independent experiments performed on three different RNA extractions from 20 to 30
male y heads. One- way ANOVA with Tukey’s post hoc test for multiple comparisons (*p<0.05, **p<0.01). (G)
Representativewestern blot of wild- type and Aprt5 mutant head extracts probed with anti- TH and anti- actin beta
antibodies. (H) Quanticationof DTH1 protein levels in adult wild- type and Aprt5 mutant heads by western blots
showed an increase in DTH1 protein level (60kDa) in Aprt5 mutants. Actin (Act5C, 42kDa) was used as a loading
control. Results are the mean of four determinations in two independent experiments. Unpaired Student’s t- test
(**p<0.01).
The online version of this article includes the following source data for gure 5:
Source data 1. Source data for Figure5A–F and H.
Source data 2. Original les for the western blot analysis of Figure5G.
Source data 3. Original les for the western blot analysis of Figure5G with relevant bands and samples labeled.
Figure 5 continued
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 15 of 34
was detected in driver and UAS control flies, and an 80.5% lower activity was detected in Drosophila
expressing the mutant form HGPRT- I42T (Table2).
We next analyzed the consequences of human HGPRT expression on the SING and BS behav-
iors. Interestingly, the pan- neuronal expression of mutant I42T isoform specifically induced a signif-
icant early locomotor defect at 15 d a.E. (PI = 0.71 vs 0.92 and 0.90 for the driver and effector
controls, respectively, p<0.01) (Figure8C) and a relatively small but robust BS behavior at 30 d a.E.
(2.3s average recovery time vs 0.64s and 0.31s for the driver and effector only controls, p<0.001)
(Figure8D). These defective phenotypes could not be seen when HGPRT- WT was expressed. There-
fore, and remarkably, whereas wild- type HGPRT expression appears to be innocuous in Drosophila,
we observed that the neuronal expression of a pathogenic LND- associated isoform triggered neuro-
behavioral impairments comparable to those of Aprt- deficient flies.
Figure 6. Relations between Aprt and molecular components of adenosinergic signaling. (A, B) Impactsof the lack of Aprt activity on the adenosinergic
system. (A)Adenosine level was measured in whole ies or heads of Aprt5 ies by ultra performance liquid chromatography (UPLC). Compared to wild-
type ies, adenosine level was signicantly reduced in the mutants. Results of three independent experiments performed with vemales per genotype
in triplicates and two independent experiments with 30 heads per genotype in triplicates. Two- way ANOVA with Sidak’s post hoc test for multiple
comparisons (*p<0.05; **p<0.01). (B)Aprt5 mutation did not affect AdoR expression but induced a marked increase in adenosine transporter Ent2
mRNA abundance. 3–6 different RNA extractions were performed on 20–30 male heads. Results of 3–6 independent experiments. Two- way ANOVA with
Sidak’s post hoc test for multiple comparisons (***p<0.001; ns: not signicant). (C)Null AdoRKGex mutants showed decreased Aprt expression (left panel)
and a stronger decrease in Aprt activity (right panel). Four independent RNA extractions were carried out on 20–30 male heads and four independent
real- time PCR determinations were done per RNA sample. For Aprt activity, three independent determinations were performed on 20 whole ies per
genotype. Unpaired Student’s t- test (***p<0.001).
The online version of this article includes the following source data for gure 6:
Source data 1. Source data for Figure6A–C.
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 16 of 34
Figure 7. Aprt deciency triggers a seizure- like phenotype. (A)At 30 days after eclosion (d a.E.), Aprt5 mutants need a much longer time than wild- type
ies to recover from a strong mechanical shock, showing a bang- sensitive (BS) paralysis comparable to tonic- clonic seizure. Results of three independent
experiments performed on 50 ies per genotype. Unpaired Student’s t- test; **p<0.01. (B)At 30 d a. E., hemizygous Aprt5 mutants also showed a marked
BS phenotype. Results of 2–4 independent experiments performed on 50 ies per genotype. One- way ANOVA with Dunnett’s post hoc test for multiple
Figure 7 continued on next page
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 17 of 34
Discussion
Over the past 35years, several animal models of LND have been developed in rodents based on
HGPRT mutation in order to better understand the cause of the disease and test potential therapeutic
treatments (Finger etal., 1988; Dunnett etal., 1989; Jinnah etal., 1993; Engle etal., 1996; Meek
et al., 2016; Witteveen et al., 2022). However, none of these models recapitulated the full LND
syndrome and, particularly, the motor or neurobehavioral symptoms resulting from HGPRT deficiency.
To date, the causes of the neurobehavioral impairments in LND are not yet clearly elucidated and
the disease is still incurable (Fu etal., 2014; Bell etal., 2016; López etal., 2020; Bell etal., 2021;
Witteveen etal., 2022). Here we used two different strategies to develop new models of this disease
in Drosophila, a useful organism to conduct genetic and pharmacological studies. First, we show that
Aprt deficiency induces both metabolic and neurobehavioral disturbances in Drosophila, similar to the
loss of HGPRT, but not APRT, activity in humans. Secondly, we expressed an LND- associated mutant
form of human HGPRT in Drosophila neurons, which also yielded behavioral symptoms. Our results
suggest that the fruit fly can be used to study the consequences of defective purine recycling pathway
and HGPRT mutation in the nervous system.
Aprt-deficient flies replicate lifespan and metabolic defects caused by
HGPRT deficiency
Flies that carry a homozygous null- mutation in Aprt develop normally and live until the adult stage
(Johnson and Friedman, 1983). While a previous study reported that heterozygous Aprt/+ flies have
an extended lifespan (Stenesen etal., 2013), we observed that homozygous Aprt5 mutants have in
contrast a significantly reduced longevity. The lack of HGPRT activity in LND also reduces lifespan
expectancy, generally under 40years of age for properly cared patients. Stenesen et al., 2013
showed that, in Drosophila, dietary supplementation with adenine, the Aprt substrate, prevented
the longevity extension conferred either by dietary restriction or heterozygous mutations of AMP
biosynthetic enzymes. This suggests that lifespan depends on accurate adenine level regulation. It
is possible that adenine could accumulate to toxic levels during aging in homozygous Aprt mutants,
explaining their shorter lifespan. Alternatively, since AMP is the Aprt product, AMP- activated protein
kinase (AMPK), an enzyme that protects cells from stresses inducing ATP depletion, could be less acti-
vated in Aprt mutants. Multiple publications explored the role of AMPK in lifespan regulation (Sinnett
and Brenman, 2016) and downregulating AMPK by RNAi in adult fat body or muscles (Stenesen
comparisons (***p<0.001). (C)RNAi- mediated downregulation of Aprt in all cells (da>AprtRNAi) also led to BS phenotype in adults at 30 d a.E., but not
with the driver and effector controls. Results of three independent experiments performed on 50 ies per genotype. One- way ANOVA with Tukey’s post
hoc test for multiple comparisons (*p<0.05). (D)Aprt knockdown by RNAi during the adult stage for 3d before the test was not sufcient to induce bang
sensitivity, suggesting that this phenotype could be caused by a developmental defect or a longer downregulation of Aprt. Results of two independent
experiments performed on 50 ies per genotype. Two- way ANOVA with Sidak’s post hoc test for multiple comparisons; ns: not signicant. (E, F)The BS
phenotype of 30- day- old Aprt5 mutants was rescued by feeding either 500µM adenosine (ado) (E)or 500µM N6- methyladenosine (m6A) (F)during all
developmental stages plus 5d before the test. Results of four or six independent experiments performed on 50 ies per genotype. One- way ANOVA
with Tukey’s post hoc test for multiple comparisons (***p<0.001, ns: not signicant).
The online version of this article includes the following video, source data, and gure supplement(s) for gure 7:
Source data 1. Source data for Figure7A–F.
Figure supplement 1. Aprt knockdown selectively in neurons, glia, or muscle cells did not induce bang sensitivity.
Figure supplement 1—source data 1. Source data for Figure7—figure supplement 1.
Figure supplement 2. Administration of allopurinol does not rescue the bang sensitivity phenotype of Aprt- decient mutants.
Figure supplement 2—source data 1. Source data for Figure7—figure supplement 2.
Figure supplement 3. Administration of various purine compounds does not rescue the motricity defects of Aprt- decient mutants.
Figure supplement 3—source data 1. Source data for Figure7—figure supplement 3.
Figure 7—video 1. Bang sensitivity phenotype of Aprt- decient ies: wild- type ies.
https://elifesciences.org/articles/88510/gures#g7video1
Figure 7—video 2. Bang sensitivity phenotype of Aprt- decient ies: Aprt5 mutant ies.
https://elifesciences.org/articles/88510/gures#g7video2
Figure 7 continued
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 18 of 34
et al., 2013), as well as its ubiquitous inactivation under starvation (Johnson et al., 2010), both
reduced fly lifespan.
In humans, HGPRT deficiency induces hypoxanthine and guanine accumulation, resulting from lack
of recycling, and increased de novo purine synthesis (Harkness etal., 1988; Fu etal., 2015; Ceballos-
Picot etal., 2015). This in turn leads to uric acid overproduction that increases the risk for nephroli-
thiasis, renal failure, and gouty arthritis if not properly treated. In insects, the end product of purine
catabolism is not uric acid but allantoin (Figure1—figure supplement 1, step 20). However, urate
oxidase, the enzyme that converts uric acid into allantoin, is specifically expressed in the Malpighi
da/+
da>HPRT1-WT
da>HPRT1-I42T
HPRT1-WT/+
HPRT1-I42T/+
0
2
4
6
8
10
HPRT1/rp49 transcript level
elav/+
elav>HPRT1-WT
HPRT1-WT/+
elav>HPRT1-I42T
HPRT1-I42T/+
0.0
0.5
1.0
Performance Index (PI)
ns
**
elav/+
elav>HPRT1-WT
HPRT1-WT/+
elav>HPRT1-I42T
HPRT1-I42T/+
0
1
2
3
Average recovery time (sec)
******
ns
Figure 8. Expression of a pathogenic mutant isoform of human hypoxanthine- guanine phosphoribosyltransferase (HGPRT) induces neurobehavioral
defects in ies. (A, B) Ubiquitousexpression of human HPRT1with da- Gal4. (A)Amplication of human HPRT1 transcripts detected by RT- PCR in head
extracts of da>HPRT1WT and da>HPRT1- I42T ies. A band with lower intensity was also detected in the effector controls (UAS- HPRT1- WT/+ and
UAS- HPRT1- I42T/+), and not in the driver control (da/+), which indicates a small amount of driver- independent transgene expression. (B)Quantication
of the previous experiment. (C, D)Expression of the Lesch–Nyhan disease (LND)- associated I42T isoform in all neurons (elav>HPRT1- I42T), but not of
the wild- type form (elav>HPRT1- WT), induced an early SING defect at 15 d a.E. (C)and a BS phenotype at 30 d a.E. (D),compared to the driver (elav/+)
and effector (UAS- HPRT1- I42T/+) only controls. Results of three independent experiments performed on 50 ies per genotype. One- way ANOVA with
Tukey’s post hoc test for multiple comparisons (**p<0.01; ***p<0.001; ns: not signicant).
The online version of this article includes the following source data for gure 8:
Source data 1. Original le for the DNA gel electrophoresis of Figure8A.
Source data 2. Original le for the DNA gel electrophoresis of Figure8A with relevant bands and samples labeled.
Source data 3. Source data for Figure8B–D.
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 19 of 34
tubules, an excretory organ producing pre- urine
(Wallrath et al., 1990). Uric acid could there-
fore accumulate in fly tissues and hemolymph if
purine salvage pathway is impaired. Accordingly,
we observed an increase in uric acid levels in
Drosophila Aprt mutant heads, which could be
reduced to normal levels by providing allopu-
rinol, a xanthine oxidase inhibitor used to protect
against renal failure in LND patients.
Aprt is required in dopaminergic
and mushroom body neurons for
young fly motricity
We found that Aprt- null adult flies show reduced
performance in the SING test, which monitors
locomotor reactivity to startle and climbing
ability. This phenotype appears at an early age,
starting from 1 d a.E. The performance continued
to decrease until 8 d a.E. and then appeared to
stabilize up to 30 d a.E. This defect is quite different from the locomotor impairments described in
Drosophila Parkinson’s disease models, in which SING performance starts declining at around 25 d
a.E. (Feany and Bender, 2000; Riemensperger etal., 2013; Rahmani etal., 2022). This phenotype
of Aprt- deficient flies could be reminiscent of the early onset of motor symptoms in LND patients,
which appear most often between 3 and 6months of age (Jinnah etal., 2006). Interestingly, knocking
down Aprt during 3d only in adult flies also induced SING impairment, which argues against a devel-
opmental flaw. Although Aprt mutants walked slower than wild- type flies, they were no less active
and their ATP levels were not different compared to controls, excluding a major failure in energy
metabolism.
Downregulating Aprt in all neurons reproduced the locomotor defect of the Aprt5 mutant, and
Aprt mutant locomotion could be partially rescued by neuronal Aprt re- expression. The fact that
rescue was not complete could be due to a dominant negative effect of the mutation, as suggested
by enzymatic assays on extracts of heterozygous Aprt5 mutants (Figure1—figure supplement 4).
Previous work from our and other laboratories showed that DA neurons control the SING behavior
in Drosophila (Feany and Bender, 2000; Friggi- Grelin etal., 2003; Riemensperger et al., 2011;
Vaccaro etal., 2017; Sun etal., 2018) and that PAM DA neurondegeneration induces SING defects
in various Parkinson’s disease models (Riemensperger etal., 2013; Bou Dib etal., 2014; Tas etal.,
2018; Pütz etal., 2021). Here, we found that knocking down Aprt either in all DA neurons or only
in the PAM DA neurons was sufficient to induce early SING defects. In contrast, knocking down Aprt
with TH- Gal4 that labels all DA neurons except for a major part of the PAM cluster did not induce
SING defects, indicating that Aprt expression in subsets of PAM neurons is critical for this locomotor
behavior.
Inactivating Aprt in all mushroom body neurons also induced a lower performance in the SING
assay. This important brain structure receives connections from DA neurons, including the PAM, and
is enriched in DA receptors (Waddell, 2010). We recently reported that activation of MB- afferent
DA neurons decreased the SING response, an effect that requires signalization by the DA receptor
dDA1/Dop1R1 in mushroom body neurons (Sun etal., 2018). We also observed a SING defect after
knocking down Aprt either in all glial cells or more selectively in glial subpopulations expressing
the glutamate transporter Eaat1, but not in the ensheathing glia. Astrocyte- like glial cells expressing
Eaat1 extend processes forming a thick mesh- like network around and inside the entire mushroom
body neuropil (Sinakevitch et al., 2010). It could be hypothesized that SING also requires Aprt
expression in these MB- associated astrocytes. However, re- expressing Aprt with Eaat1- Gal4 did not
lead to SING rescue in Aprt mutant background. This suggests that the presence of Aprt in neurons
can somehow compensate for Aprt deficiency in glia, but the reverse is not true. In conclusion, proper
startle- induced locomotion in young flies depends on Aprt activity in PAM DA and mushroom body
neurons, and in Eaat1- expressing glial cells.
Table 2. Hypoxanthine- guanine
phosphoribosyltransferase (HGPRT) activity in
transgenic flies expressing the wild- type or a
Lesch–Nyhan disease (LND)- associated mutant
form of human HPRT1.
Genotypes HGPRT activity (nmol/min/mg)
da/+ 0
da>HPRT1- WT 13.88 ± 3.75
da>HPRT1- I42T 2.70 ± 1.44
HPRT1- WT/+ 0
HPRT1- I42T/+ 0
The online version of this article includes the following
source data for table 2:
Source data 1. Source data for Table2.
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 20 of 34
Neuronal Aprt regulates spontaneous activity and sleep
Because sleep is regulated by the mushroom body as well as DA neurons in flies (Artiushin and
Sehgal, 2017), we monitored spontaneous locomotor activity and sleep pattern of Aprt5 mutants.
Their activity profile appeared normal, with unaltered morning and evening anticipation, indicating
that the circadian rhythms are surprisingly maintained in light- dark (LD) conditions in the absence of
a functional purine recycling pathway (Figure4—figure supplement 1). This experiment also high-
lighted that Aprt- deficient flies are hyperactive during the night, which suggests that they sleep less
than wild- type flies. This could be confirmed by measuring their sleep pattern on video recordings.
Aprt mutants show a reduced walking speed, sleep less during both day and night, and have diffi-
culty in maintaining sleep. Downregulating Aprt selectively in neurons reproduced the sleep defect,
whereas doing it in glial cells only had no effect. We have not attempted here to identify further the
neuronal cells involved in this phenotype. Like for the SING behavior defect, it could involve PAM DA
neurons as subpopulations of this cluster have been shown to regulate sleep in Drosophila (Nall etal.,
2016). Therefore, the lack of purine recycling markedly disrupts sleep in Aprt- deficient flies, making
them more active at night. This is strikingly comparable to young LND patients who have a much
disturbed sleep time during the night (Mizuno etal., 1979).
Lack of Aprt reduces adenosine signaling leading to DA neuron
overactivation
We observed that Aprt deficiency did not decrease DA levels in the Drosophila brain. This prompted
us to study another neuromodulator, adenosine, which is indirectly a product of Aprt enzymatic
activity. The purine nucleoside adenosine is one of the building blocks of RNA and the precursor of
ATP and cAMP, but is also the endogenous ligand of adenosine receptors that modulate a wide range
of physiological functions. In brain, adenosine regulates motor and cognitive processes, such as the
sleep- wake cycle, anxiety, depression, epilepsy, memory, and drug addiction (Soliman etal., 2018).
In normal metabolic conditions, adenosine is present at low concentrations in the extracellular space
and its level is highly regulated, either taken up by cells and incorporated into ATP stores or deami-
nated by adenosine deaminase into inosine. In mammals, several nucleoside transporters mediate the
uptake of adenosine and other nucleosides into cells, named equilibrative and concentrative nucleo-
side transporters, respectively (Gray etal., 2004; Boswell- Casteel and Hays, 2017; Pastor- Anglada
and Pérez- Torras, 2018).
We observed a marked reduction in adenosine levels in Aprt mutant flies. While we did not observe
any alteration in the transcript levels of AdoR, the gene coding for the only adenosine receptor in
Drosophila, transcript levels of an adenosine transporter, Ent2, were increased more than twofold in
Aprt mutants. Interestingly, one paper reported a similar strong increase in Ent2 expression in AdoR
mutant flies (Knight etal., 2010). These results suggest that AdoR signaling is less activated in Aprt-
deficient flies compared to controls. We also observed that the lack of AdoR decreased Aprt expres-
sion and activity in Drosophila, possibly from increased Ent2 expression and so higher adenosine influx
which could downregulate Aprt by a feedback mechanism.
Adenosine and DA receptors are known to interact closely in mammals (Franco etal., 2000; Kim
and Palmiter, 2008). A previous study performed in Drosophila larvae showed that an increase in
astrocytic Ca2+ signaling can silence DA neurons through AdoR stimulation by a mechanism potentially
involving the breakdown of released astrocytic ATP into adenosine (Nall etal., 2016). We previously
showed that DA neuron activation can decrease fly performance in the SING test (Sun etal., 2018).
Fly nocturnal hyperactivity can also be caused by an increase in DA signaling (Kumar etal., 2012;
Lee etal., 2013), in accordance with our observation that Aprt- deficient flies sleep less and are more
active during the night. Therefore, both the locomotor and sleep defects induced by the lack of Aprt
activity could be explained by DA neuron overactivation that would result from reduced adenosinergic
signaling.
Adenosine has been proposed before to be involved in neurological consequences of LND (Visser
et al., 2000). Adenosine transport is decreased in peripheral blood lymphocytes of LND patients
(Torres etal., 2004), as well as A2A adenosine receptor mRNA and protein levels (García etal., 2009;
García etal., 2012). In a murine LND model, adenosine A1 receptor expression was found to be
strongly increased and that of A2A slightly decreased, while A2B expression was not affected (Bertelli
et al., 2006). Chronic administration of high doses of caffeine, an adenosine receptor antagonist,
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 21 of 34
caused self- injurious behavior in rats (Miñana etal., 1984; Jinnah etal., 1990). Moreover, central
injection of an adenosine agonist is sufficient to prevent self- mutilation induced by dopaminergic
agonist administration in neonatally 6- OHDA lesioned rats (Criswell etal., 1988).
Adenosine or N6-methyladenosine supplementation rescues the
epilepsy behavior of Aprt mutants
LND is characterized by severe behavioral troubles, including dystonia, spasticity, and involuntary
movements (choreoathetosis). Some patients can also show an epileptic disorder (Jinnah etal., 2006;
Madeo etal., 2019). In flies, the BS test is often used to model epileptic seizures (Song and Tanouye,
2008; Parker et al., 2011). Here, we observed that 30- day- old Aprt mutant flies show a transient
seizure- like phenotype after a strong mechanical shock. Although seizure duration appeared shorter
in Aprt5 than in typical BS mutants such as bss, at least one other BS mutant, porin, was reported
to have similar short recovery times as Aprt- deficient flies (Graham et al., 2010). Previous works
demonstrated that BS is linked to neuronal dysfunction in Drosophila (Parker etal., 2011; Kroll and
Tanouye, 2013; Kroll etal., 2015; Saras and Tanouye, 2016). Knocking down Aprt in specific cells
such as neurons, glia, or muscles did not trigger this phenotype, suggesting that Aprt must be inacti-
vated in several cell types to induce seizure.
Interestingly, knocking down Aprt by RNAi in all cells during development also induced the BS
behavior, but not for 3d only in adult flies, at variance with the SING phenotype. We have fed the
mutants with a diet supplemented with various compounds involved in purine metabolism, including
allopurinol, adenine, hypoxanthine, adenosine, or its analog N6- methyladenosine (m6A) either
throughout larval development or in adult stage. Only adenosine and m6A, ingested during devel-
opment, rescued the BS behavior. This suggests that loss of Aprt induces a lack of adenosine in the
developing nervous system, as we observed in adult flies (Figure6), which may alter neural circuits
controlling BS behavior in adults. The adenosine analog m6A cannot be incorporated into nucleic acids
and is excreted in the urine (Schram, 1998; Batista, 2017). The rescuing effect we observed with
m6A suggests thereby that both this compound and adenosine are required as adenosine receptor
agonists or allosteric regulators during development, rather than nucleotide precursors.
Adenosine can strongly inhibit cerebral activity and its role as endogenous anticonvulsant and
seizure terminator is well established in humans (Boison, 2005; Masino etal., 2014; Weltha etal.,
2019). Conversely, deficiencies in the adenosine- based neuromodulatory system can contribute
to epileptogenesis. For instance, increased expression of astroglial adenosine kinase (ADK), which
converts adenosine into AMP, leads to a reduction in brain adenosine level that plays a major role in
epileptogenesis (Weltha etal., 2019). Hence, therapeutic adenosine increase is a rational approach
for seizure control. Our observation that feeding adenosine or its derivative m6A rescued the seizure-
like phenotype of Aprt mutant flies further suggests that adenosinergic signaling has partly similar
functions in the fly and mammalian brains. In addition, the decrease in adenosine levels we observed
in Aprt mutants could result from enhanced ADK activity that would compensate for the lack of Aprt-
produced AMP.
We and others recently observed that m6A and related compounds sharing an adenosine moiety
are able to rescue the viability of LND fibroblasts and neural stem cells derived from induced plurip-
otent stem cells (iPSCs) of LND patients cultured in the presence of azaserine, a potent blocker of de
novo purine synthesis (Petitgas and Ceballos- Picot, unpublished results; Ruillier etal., 2020). Like in
flies again, allopurinol was not capable of rescuing the cell viability in this in vitro model. The similarity
of these results increases confidence that Aprt- deficient Drosophila could be used as an animal model
of LND.
Expression of mutant HGPRT triggers locomotor and seizure
phenotypes
How HGPRT deficiency can cause such dramatic neurobehavioral troubles in LND patients still remains
a crucial question. To date, cellular (Smith and Friedmann, 2000; Torres etal., 2004; Ceballos- Picot
etal., 2009; Cristini etal., 2010; Guibinga etal., 2012; Sutcliffe etal., 2021) and rodent (Finger
etal., 1988; Dunnett etal., 1989; Jinnah etal., 1993; Meek etal., 2016; Witteveen etal., 2022)
models only focused on the consequences of HGPRT deficiency to phenocopy the disease. Such an
approach was justified by the fact that the lower the residual activity of mutant HGPRT, the more
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 22 of 34
severe the symptoms are in patients (Fu and Jinnah, 2012. Fu et al., 2014). However, it could be
conceivable that part of these symptoms result from compensatory physiological mechanisms or a
deleterious gain- of- function conferred to the HGPRT protein by the pathogenic mutations. Here, we
observed that neuronal expression of mutant human HGPRT- I42T, which expresses a low enzymatic
activity, but not the wild- type HGPRT protein, induced early locomotor defects in young flies and BS in
older flies, similarly to the defects induced by Aprt deficiency. This suggests a potential neurotoxicity
of the pathological mutant form of HGPRT, which could be related to disturbances in purine metab-
olism or other signaling pathways. The human mutant form might not be properly degraded and
accumulate as aggregates, potentially exerting neuronal toxicity. Such an approach opens interesting
perspectives to better understand the endogenous function of HGPRT and its pathogenic forms.
Indeed, the identification of a potential inherent neurotoxicity of defective forms of human HGPRT is
a new element, which could be explored in further work in the fly and also in rodent models.
A new model of LND in an invertebrate organism?
LND, a rare X- linked metabolic disorder due to mutations of the HPRT1 gene, has dramatic neuro-
logical consequences for affected children. To date, no treatment is available to abrogate these trou-
bles, and no fully satisfactory in vivo models exist to progress in the understanding and cure of this
disease. HGPRT knockout rodents do not show comparable motor and behavioral troubles, which
makes these models problematic for testing new therapeutic treatments. Drosophila does not express
HGPRT- like activity and our phylogenetic analysis established that no HGPRT homolog is present in
D. melanogaster (Figure1—figure supplement 2 and Figure1—figure supplement 3), confirming
that Aprt is the only enzyme of the purine salvage pathway in this organism. APRT and HGPRT are
known to be functionally and structurally related. Both human APRT and HGPRT belong to the type I
PRTases family identified by a conserved phosphoribosyl pyrophosphate (PRPP) binding motif, which
is used as a substrate to transfer phosphoribosyl to purines. This binding motif is only found in PRTases
from the nucleotide synthesis and salvage pathways (Sinha and Smith, 2001). Moreover, the purine
substrates adenine, hypoxanthine, and guanine share the same chemical skeleton and APRT can bind
hypoxanthine, indicating that APRT and HGPRT also share similarities in their substrate binding sites
(Ozeir etal., 2019).
Here, we find that Aprt mutant flies show symptoms partly comparable to the lack of HGPRT in
humans, including increase in uric acid levels, reduced longevity, and various neurobehavioral defects
such as early locomotor decline, sleep disorders, and epilepsy behavior. This animal model therefore
recapitulates both salvage pathway disruption and motor symptoms, as observed in LND patients.
Moreover, our results highlight that Aprt deficiency in Drosophila has more similarities with HGPRT
than APRT deficiency in humans. Aprt mutant flies also show a disruption of adenosine signaling, and
we found that adenosine itself or a derivative compound can relieve their epileptic symptoms. Finally,
neuronal expression of a mutant form of human HGPRT that causes LND also triggers abnormalities
in fly locomotion and seizure- like behavior, which has not been documented to date in other models.
The use of Drosophila to study LND could therefore prove valuable to better understand the link
between purine recycling deficiency and brain functioning and carry out drug screening in a living
organism, paving the way toward new improvements in curing this dramatic disease.
Materials and methods
Drosophila culture and strains
Flies were raised at 25°C on standard cornmeal- yeast- agar medium supplemented with methyl hydroxy-
benzoate as an antifungal under a 12hr:12hr LD cycle. The Drosophila mutant lines were obtained
either from the Bloomington Drosophila Stock Center (BDSC), the Vienna Drosophila Resource Center
(VDRC) or our own collection: Aprt5 (Johnson and Friedman, 1983) (BDSC #6882), Df(3L)ED4284
(BDSC #8056), Df(3L)BSC365 (BDSC #24389), UAS- AprtRNAi (VDRC #106836), AdoRKG03964ex (Wu etal.,
2009) here named AdoRKGex (BDSC #30868), and the following Gal4 drivers: 238Y- Gal4, 24B- Gal4, da-
Gal4, Eaat1- Gal4 (Rival etal., 2004), elav- Gal4 (elavC155), MZ0709- Gal4, NP6510- Gal4, NP6520- Gal4,
R58E02- Gal4 (Liu etal., 2012), R76F05- Gal4, repo- Gal4, TH- Gal4 (Friggi- Grelin etal., 2003), TRH-
Gal4 (Cassar et al., 2015), tub- Gal80ts , and VT30559- Gal4. The Canton- S line was used as wild-
type control. The simplified driver>effector convention was used to indicate genotypes in figures, for
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 23 of 34
example, elav>AprtRNAi for elav- Gal4; UAS- AprtRNAi. In some experiments, to restrict RNAi- mediated
Aprt inactivation at the adult stage, we have used the TARGET system (McGuire etal., 2003). Flies
were raised at 18°C (permissive temperature) where Gal4 transcriptional activity was inhibited by
tub- Gal80ts, and shifted to 30°C (restrictive temperature) for 3d before the test to enable Gal4- driven
AprtRNAi expression.
Construction of transgenic lines
To generate a UAS- Aprt strain, a 549bp Aprt insert containing the coding sequence was PCR amplified
from the ORF clone BS15201 (Drosophila Genomics Resource Center, Bloomington, IN) using primers
with added restriction sites (in bold type): forward 5- AGGG AATT GG GAAT TC GTTA TCAG TCGA
CATG AGCC C, reverse 5- ACAA AGAT CC TCTA GA TCTA GAAA GCTT TCAG TACT TAAT G. After diges-
tion with EcoRI and XbaI, the Aprt cDNA was subcloned into the pUASTattB vector (Bischof etal.,
2007) using the In- Fusion HD Cloning Kit (Takara Bio, Kyoto, Japan) according to the manufacturer’s
instructions, and the insertion verified by sequencing (Eurofins Genomics, Ebersberg, Germany). The
construction was sent to BestGene (Chino Hills, CA, USA) for Drosophila germline transformation into
the attP14 docking site on the 2d chromosome. The UAS- Aprt line yielding the strongest expression
was selected and used in the experiments.
A clone containing the human wild- type HPRT1 cDNA was kindly provided to us by Prof. Hyder
A. Jinnah (Emory University, GA). We constructed the HPRT1- I42T cDNA from this clone by site-
directed mutagenesis using the QuikChange II Site- Directed Mutagenesis Kit (Agilent Technologies,
Santa Clara, CA). Primers sequences used to create the mutation were: forward 5'- CAGT CCTG TCCA
TA G TTAG TCCA TGAG GAAT AAAC ACCC T and reverse 5'- AGGG TGTT TATT CCTC ATGG ACTA A C TATG
GACA GGAC TG (the bases modified to create the mutation are in bold type). The cDNA obtained
was verified by sequencing. Then, the 657bp HPRT1- WT and HPRT1- I42T inserts were PCR amplified
using primers with added restriction sites (in bold type) and Drosophila translation start consensus
sequence: forward 5- AGGG AATT GG GAAT TC AAGA AGGA GAT ACAA A ATGG C and reverse 5- ACAA
AGAT CC TCTA GA GCTC GGAT CCTT ATCA TTAC (the bases modified to match the Drosophila transla-
tion initiation consensus sequence are underlined). They were subcloned into pUASTattB and verified
by sequencing. The transgenes were sent to BestGene for Drosophila transformation and inserted
into the attP40 docking site on the 2d chromosome.
Sequencing of Aprt5
For sequencing of the Aprt5 cDNA, total RNA was isolated by standard procedure from 20 to 30 heads
of homozygous Aprt5 flies and reverse transcribed using oligo d(T) primers (PrimeScript RT Reagent
Kit, Takara Bio). At least 750ng of the first- strand cDNA was amplified by PCR in 20μl of reaction
mixture using PrimeStar Max DNA polymerase (Takara Bio) with a Techne Prime Thermal Cycler appa-
ratus (Bibby Scientific, Burlington, NJ). The program cycles included 10s denaturation at 98°C, 10s
annealing at 55°C, and 10s elongation at 72°C, repeated 35 times. 1μl of the PCR product was
amplified again with the same program, in 30μl of reaction mixture. After elution on 1% agarose gel,
DNA were purified using the Wizard SV Gel and PCR Clean- Up System protocol (Promega, Madison,
WI) according to the manufacturer’s instructions. Finally, 7.5μl of purified DNA were sent with 2.5μl
of primers (forward and reverse in separate tubes) for sequencing (Eurofins Genomics).
Phylogenetic analyses
HGPRT homologs were identified by BlastP searching the NCBI GenBank non- redundant protein
database (last October 2019 version). A subset of interest was selected for phylogenetic analyses. For
Figure1—figure supplement 2, multiple sequence alignment was performed using MAFFT (- ensi)
(Katoh and Standley, 2013). The confidence of aligned residues was assessed using the TCSindex
(Chang etal., 2014) only columns with TCS index 7–9 (on a 0–9 scale) were retained. ProtTest v3.2
(Darriba etal., 2011) was used to assess the best model fitting of the data. Maximum likelihood tree
was inferred in IQ- TREE (Trifinopoulos etal., 2016), under the LG + R6 model. Bayesian inference
was carried out in PhyloBayes v. 3.3 (Lartillot etal., 2009), under the LG + Γ4 model. For Figure1—
figure supplement 3, the whole analysis was performed in SeaView 5.0.4 (Gouy et al., 2010).
Multiple sequence alignment was performed using MUSCLE (Edgar, 2004). Maximum likelihood tree
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 24 of 34
was inferred using PhyML 3.0 (Guindon etal., 2010) (under the LG + Γ4 model; best of NNI and SPR
tree searching option; invariable sites optimized).
Lifespan assay
Longevity study was performed as previously described (Riemensperger etal., 2011). Drosophila
adult males were collected within 24hr of emergence and maintained on standard medium at 25°C
under a 12:12hr LD cycle. They were transferred into fresh bottles every 2–3d, and the number or
surviving flies was scored. Also, 50 flies per bottle in triplicate were tested for each genotype and the
experiment was done three times.
Startle-induced negative geotaxis (SING)
SING assays were performed as previously described (Rival etal., 2004; Riemensperger etal., 2011;
Sun etal., 2018). For each genotype, 50 adult males divided into five groups of 10 flies were placed in
a vertical column (25cm long, 1.5cm diameter) with a conic bottom end and left for about 30min for
habituation. Then, columns were tested individually by gently tapping them down (startle), to which
flies normally responded by climbing up. After 1min, flies having reached at least once the top of the
column (above 22cm) and flies that never left the bottom (below 4cm) were counted. Each fly group
was assayed three times at 15min intervals. The PI for each column is defined as ½[1 + (ntop- nbot)/ntot],
where ntot is the total number of flies in the column, ntop the number of flies at the top, and nbot the
number of flies at the bottom after 1min. SING was performed at 1, 8, 10, and 30d a.E.
Spontaneous locomotion and sleep monitoring
Spontaneous locomotor activity was recorded as previously described (Vaccaro etal., 2017) using
Drosophila activity infrared beam monitors (DAM, TriKinetics Inc, Waltham, MA) placed in incubators
at 25°C equipped with standard white light. Eight- day- old male flies were maintained individually for
5–6d under 12:12hr LD cycle in 5 × 65mm glass tubes containing 5% sucrose, 1.5% agar medium.
Data analysis was performed with the FaasX software (Klarsfeld etal., 2003). Histograms represent
the distribution of the activity through 24hr in 30min bins, averaged for 32 flies per genotype over
4–5 cycles.
For sleep monitoring, 2–4- day- old virgin female flies were transferred individually into 5 × 65mm
glass tubes containing standard food and their movements were recorded for up to 5d using DAM
infrared beam monitors (TriKinetics Inc) or the Drosophila ARousal Tracking (DART) video system
(Faville etal., 2015), in a 12hr:12hr LD cycle with 50–60%humidity. Control and experimental flies
were recorded simultaneously. Each experiment included at least 14 flies for each condition and was
repeated 2–3 times with independent groups of flies. Fly sleep, defined as periods of immobility
lasting more than 5min (Huber et al., 2004), was computed with a Microsoft Excel macro for the
infrared beam data and with the DART software for the video data. Distribution and homogeneity as
well as statistical group comparisons were tested using Microsoft Excel plugin software StatEL. The
p- value shown is the highest obtained among post hoc comparisons and means ± SEM were plotted.
Immunohistochemistry
Brains of 8–10- day- old adult flies were dissected in ice- cold Ca2+- free Drosophila Ringer’s solu-
tion and processed for whole- mount anti- TH or anti- DA immunostaining as previously described
(Riemensperger etal., 2011; Cichewicz et al., 2017). The primary antibodies used were mouse
monoclonal anti- TH (1:1000, Cat# 22941, ImmunoStar, Hudson, WI) and mouse monoclonal anti- DA
(1:100, Cat# AM001, GemacBio, Saint- Jean- d'Illac, France). The secondary antibody was goat anti-
mouse conjugated to Alexa Fluor 488 (1:1000, Thermo Fisher Scientific, Waltham, MA). Brains were
mounted with antifade reagent, either Prolong Gold (Thermo Fisher Scientific) or, alternatively, 65%
2,2-thiodiethanol (Sigma- Aldrich, St. Louis, MI) (Cichewicz et al., 2017) for DA staining. Images
were acquired on a Nikon A1R confocal microscope with identical laser, filter, and gain settings
for all samples in each experiment. Immunofluorescence intensity levels were quantified using the
Fiji software. Experiments were repeated independently at least three times on 4–6 brains per
genotype.
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 25 of 34
Bang sensitivity test
Bang sensitivity assays were performed as previously described (Howlett etal., 2013). 30- day- old
males were divided into five groups of 10 flies under CO2 and allowed to recover overnight. The
following day, each group was placed in a vial without food and after 20min of habituation, the vials
were stimulated vigorously with a vortex mixer for 10s at 2500rpm. The recovery time was measured
for each fly, from the end of the stimulation until they reached a normal standing position. Results are
the mean of the recovery time for at least 50 flies per genotype.
Drug administration
Allopurinol (A8003, Sigma- Aldrich) was diluted in standard medium at 100μg/ml and flies were placed
for 5d on this medium before metabolite extraction. Adenine, adenosine, and hypoxanthine (A2786,
A9251, and H9377, Sigma- Aldrich) or N6- methyladenosine (m6A) (QB- 1055, Combi- Blocks, San Diego,
CA) were diluted in fly food medium at 500μM. Parents were allowed to lay eggs on this medium in
order to have exposition to the drug throughout all larval development of the progeny. Adults were
collected and placed in normal medium until 5d before the test, when they were placed again in food
supplemented with adenosine or m6A at the same concentrations.
Uric acid quantification
For purine metabolite extraction, 40 heads from 8- day- old flies were ground in 80% ethanol, heated
for 3min at 80°C, and evaporated in a Speedvac apparatus. Dried residues were resuspended in MilliQ
water and total protein content of each homogenate was measured with the Pierce BCA Protein Assay
Kit (Thermo Fisher Scientific). 20µl of each sample was injected into a 25cm × 4.6mm C18 Nucleosil
column with 5 µm particles (Interchim, Montluçon, France). Purine metabolites were detected by
an Agilent 1290 Infinity HPLC system coupled with a diode array detector as recommended by the
ERDNIM advisory document. Seven wavelengths were used for detection (230, 240, 250, 260, 266,
270, and 280nm) in order to have the spectrum of each compound for identification in addition to the
retention time. The mobile phases contained 0.05M monopotassium phosphate pH 5 and 65% (v/v)
acetonitrile. The flux was fixed at 1ml/min. For analysis, the maximum height of the compound was
normalized to protein content and compared to the control genotype.
ATP assay
ATP level was measured by bioluminescence using the ATP Determination Kit (A22066, Thermo Fisher
Scientific) on a TriStar 2 Spectrum LB942S microplate reader (Berthold Technologies, Bad Wildbad,
Germany) according to the manufacturer’s instructions. Samples were prepared as described previ-
ously (Fergestad etal., 2006). Briefly, 30 heads or 5 thoraces from 8- day- old flies were homogenized
in 200µl of 6M guanidine- HCl to inhibit ATPases, boiled directly 5min at 95°C, and then centrifuged
for 5min at 13,000× g and 4°C. Total protein content of each supernatant was measured with the
BCA Protein Assay Kit. Each supernatant was then diluted at 1:500 in TE buffer pH 8 and 10µl were
placed in a 96- well white- bottom plate. The luminescent reaction solution (containing d- luciferine,
recombinant firefly luciferase, dithiothreitol, and reaction buffer) was added to each well with an
injector and high- gain 1 s exposure glow reads were obtained at 15 min after reaction initiation.
Results were compared to a standard curve generated with known ATP concentrations, and final
values were normalized to the protein content.
Enzyme activity assay
HGPRT and APRT activities were assessed in Drosophila by adapting the methods previously estab-
lished for human cells (Cartier and Hamet, 1968; Ceballos- Picot et al., 2009). Twenty whole male
flies were homogenized in 250µl of 110 mM Tris, 10mM MgCl2 pH 7.4 (Tris- MgCl2 buffer), imme-
diately frozen and kept at least one night at –80°C before the assay. After 5min of centrifugation
at 13,000× g, total protein content of each supernatant was measured with the BCA Protein Assay
Kit to normalize activity level. Kinetics of [14C]-hypoxanthine conversion to IMP (or in some cases
[14C]-guanine conversion to GMP), and [14C]-adenine conversion to AMP were assessed for HGPRT
and Aprt assays, respectively. Compositions of reaction mediums were 25µl of a radioactive solution
made with 38µl of [14C]-hypoxanthine (25 µCi/ml) diluted in 1 ml of 1.2mM cold hypoxanthine (or
in some cases 38µl of [14C]-guanine [25µCi/ml] diluted in 1ml of 1.2mM cold guanine), for HGPRT
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 26 of 34
assay, or 25µl of a radioactive solution made with 75µl of [14C]-adenine (50µCi/ml) diluted in 1ml of
1.2mM cold adenine for Aprt assay, and a volume of fly extract equivalent to 200µg protein diluted
in Tris- MgCl2 buffer to 150µl. Reactions were monitored at 37°C and started by adding 25µl of the
co- factor 5- phosphoribosyl- 1- pyrophosphate (PRPP) at 10mM. After 6, 12, 24, and 36min, 40µl of
each pool was placed in a tube containing either 25µl of HIE (3mM hypoxanthine, 6mM IMP, 200mM
EDTA) or AAE (3mM adenine, 6mM AMP, 200mM EDTA) solutions and incubated for 3min at 95°C
to stop the reaction. The different radioactive compounds were separated by paper chromatography
on Whatman 3MM strips using 28% NH4OH, 50mM EDTA as solvent for about 1 hr 30 min. Then, the
substrates and products were visualized under a UV lamp at 254nm and placed separately in vials in
2ml of Scintran (VWR, Radnor, PA). The radioactivity in disintegrations per minute was measured in
a Packard Tri- Carb 1600 TR liquid scintillation analyzer (PerkinElmer, Waltham, MA). The percentage
of substrate transformation as a function of time was converted in nmol/min/mg protein and finally
normalized to wild- type control values.
Protein extraction and western blots
Thirty heads of 8–10- day- old males per genotype were homogenized in 30 μl Laemmli buffer
containing protease inhibitor (cOmplete mini Protease Inhibitor Cocktail, Roche Diagnostics) using a
Minilys apparatus (Bertin Instruments, Montigny- le- Bretonneux, France). The lysates were incubated
on ice for 30min and centrifuged at 8000× g for 10min at 4°C. The extracted proteins were heated
at 95°C for 10min. Western blots were performed as previously described (Issa etal., 2018). Briefly,
proteins were separated in 4–12% Novex NuPAGE Bis- Tris precast polyacrylamide gels (Life Tech-
nologies) following the manufacturer’s protocol in a MOPS- SDS running buffer. A semi- dry transfer
was done onto polyvinylidene difluoride membranes (Amersham Hybond P 0.45μm) using a Hoefer
TE77 apparatus. Membranes were probed overnight at 4°C with mouse monoclonal anti- TH (1:5000,
Cat# 22941, ImmunoStar) and mouse monoclonal anti- actin beta (1:5000, Cat# ab20272, Abcam,
Cambridge, UK) that cross- reacts with Drosophila Actin 5C (Act5C). After incubation with horseradish
peroxidase (HRP)- conjugated anti- mouse (1:5000, Cat# 115- 035- 146, Jackson ImmunoResearch) as
secondary antibody, immunolabeled bands were revealed by chemiluminescence staining using ECL
RevelBlOt Intense (Ozyme, Saint- Cyr- l'École, France, Cat# OZYB002- 1000) and then digitally acquired
with the ImageQuant TL software (GE Healthcare Life Science). Densitometry measures were made
with the Fiji software and normalized to Act5C values as internal controls.
Reverse transcription-PCR and quantitative PCR
Total RNA was isolated by standard procedure from 20 to 30 heads of 8- day- old males collected on
ice and lysed in 600µl QIAzol Reagent (QIAGEN, Venlo, Netherlands). 1μg of total RNA was treated
by DNase (DNase I, RNase- free, Thermo Fisher Scientific) according to the manufacturer’s instructions.
5μl of treated RNA was reverse transcribed using oligo d(T) primers (PrimeScriptRT Reagent Kit,
Takara Bio). Then, at least 750ng of the first- strand cDNA was amplified in 20μl of reaction mixture
using PrimeStar Max DNA polymerase (Takara Bio) with a Techne Prime Thermal Cycler apparatus
(Bibby Scientific). The program cycles included 10s denaturation at 98°C, 10s annealing at 55°C, and
10s elongation at 72°C, repeated 30 times. PCR product levels were measured after electrophoresis
by densitometry analysis with the Fiji software. Data were normalized to amplification level of the
ribosomal protein rp49/RpL32 transcripts as internal control. Sequences of the primers used were
for HPRT: forward 5- GAGA TACA AAAT GGCG ACCC GCAG CCCT , reverse, 5- GCTC GGAT CCTT ATCA
TTAC TAGG CTTT G (amplicon 686bp); and for rp49: forward 5- GACG CTTC AAGG GACA GTAT C and
reverse rp49, 5AAAC GCGG TTCT GCAT GAG (amplicon 144bp).
For RT- qPCR, approximately 40 ng of the first- strand cDNA was amplified in 10 μl of reaction
mixture using the LightCycler 480 SYBR Green I Master reaction mix (Roche Applied Science,
Mannheim, Germany) and a LightCycler 480 Instrument (Roche Applied Science). The program
cycles included a 10min preincubation step at 95°C, 40 cycles of amplification (10s denaturation
at 95°C, 10 s annealing at 55°C, 20 s elongation at 72°C), followed by a melting curves analysis
for PCR product identification. Data were normalized to amplification level of the ribosomal protein
rp49/RpL32 transcripts as internal control. The genes analyzed and primer sequences used for qPCR
are AdoR, forward 5- GGAG AAAT TGCG ATCG GATG ACAC , reverse 5- TCTT CAGC GAAC TCCG AGTG
AATG ; Aprt, forward 5- AATC AGCG CGGA AGAC AAGC TA, reverse, 5- CCAC CTTG CCGA TGAG TTCA
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 27 of 34
GT; DTH1, forward 5- GGAT CGAA AGCC AACC AAGT G, reverse 5- CTTG GGGA CCAA CTGC GCTT TA;
Ent2, forward 5- ACGG CAAG GGAT CAAC GTC, reverse 5- CCGT GCAG CAGG AATA TAAA GA; rp49,
forward 5- GACG CTTC AAGG GACA GTAT C; reverse 5- AAAC GCGG TTCT GCAT GAG.
Adenosine assay
Adenosine was determined by ultra performance liquid cChromatography (UPLC). Here, 5 whole
flies or 30 heads were homogenized in 120 µl 0.9% (w/v) NaCl using a Minilys apparatus (Bertin
Instruments) and frozen at –80°C. After unfreezing and 5min of microcentrifugation, 20 µl of 10%
perchloric acid were added to 70µl of the supernatant and the mixture was left for 5min on ice. After
a new centrifugation, 20µl of a neutralization solution (made by mixing 3 volumes of 3M K2CO3 with
1volume of 6.4mM NaOH containing 0.4mg/ml bromothymol blue) were added and the mixture was
centrifuged again before injection (5µl). Samples were analyzed with a UV diode- array detector on an
Acquity UPLC HSS T3 column (1,8µm, 2.1 × 150mm) (Waters Corporation, Milford, MA). The mobile
phases consisted of Buffer A (30mM ammonium acetate, pH 4.0 with 1:10,000 heptafluorobutyric
acid [HPFA]) and Buffer B (acetonitrile with 1:10,000 HPFA) using a flow rate of 0.3ml/min. Chromato-
graphic conditions were 3.5min 100% Buffer A, 16.5min up to 6.3% Buffer B, 2min up to 100% Buffer
B, and 1min 100% Buffer B. The gradient was then returned over 5min to 100% Buffer A, restoring
the initial conditions. Results were normalized to protein levels for each sample.
Statistics
Statistical significance was determined using the Prism 6 software (GraphPad Software, La Jolla,
CA). Survival curves for longevity experiments were analyzed using the log- rank test. Student’s t- test
was used to compare two genotypes or conditions, and one- way or two- way ANOVA with Tukey’s,
Dunnett’s, or Sidak’s post hoc multiple comparison tests for three or more conditions. Results are
presented as mean ± SEM. Probability values in all figures: *p<0.05, **p<0.01, ***p<0.001.
Acknowledgements
We are grateful to Pr. Hyder A Jinnah for the gift of HPRT cDNAs and to Dr Thomas Riemensperger
for helpful discussions. Part of the phylogenetic analyses were performed using the computing facil-
ities of the PRABI- AMSB bioinformatics platform, Laboratory of Biometry and Evolutionary Biology,
Lyon, France. This work was supported by funding from CNRS and ESPCI Paris to SB and Association
Malaury to ICP. CP is a recipient of PhD fellowships from the Association Lesch- Nyhan Action (LNA)
and Labex MemoLife.
Additional information
Funding
Funder Grant reference number Author
Centre National de la
Recherche Scientique Laboratory funding Serge Birman
École Supérieure de
Physique et de Chimie
Industrielles de la Ville de
Paris
Laboratory funding Serge Birman
Association Malaury Laboratory funding Irène Ceballos-Picot
Association Lesch-Nyhan
Action Graduate Student
Fellowship Céline Petitgas
Labex MemoLife Graduate Student
Fellowship Céline Petitgas
The funders had no role in study design, data collection and interpretation, or the
decision to submit the work for publication.
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 28 of 34
Author contributions
Céline Petitgas, Conceptualization, Formal analysis, Validation, Investigation, Methodology, Writing
– original draft, Writing – review and editing; Laurent Seugnet, Sandrine Marie, Conceptualization,
Formal analysis, Investigation, Methodology; Amina Dulac, Rebecca Fima, Marion Strehaiano, Joana
Dagorret, Formal analysis, Investigation; Giorgio Matassi, Formal analysis, Investigation, Method-
ology; Ali Mteyrek, Baya Chérif- Zahar, Methodology; Irène Ceballos- Picot, Conceptualization, Super-
vision, Funding acquisition, Validation, Writing – original draft, Writing – review and editing; Serge
Birman, Conceptualization, Supervision, Funding acquisition, Validation, Writing – original draft,
Project administration, Writing – review and editing
Author ORCIDs
Laurent Seugnet
http://orcid.org/0000-0003-1617-5721
Giorgio Matassi
http://orcid.org/0000-0003-4923-226X
Serge Birman
http://orcid.org/0000-0002-4278-454X
Peer review material
Reviewer #1 (Public review): https://doi.org/10.7554/eLife.88510.3.sa1
Reviewer #2 (Public review): https://doi.org/10.7554/eLife.88510.3.sa2
Reviewer #3 (Public review): https://doi.org/10.7554/eLife.88510.3.sa3
Author response https://doi.org/10.7554/eLife.88510.3.sa4
Additional files
Supplementary files
MDAR checklist
Data availability
All data generated or analysed during this study are included in the manuscript and supporting files;
source data files have been provided for all the figures and linked figure supplements.
References
Artiushin G, Sehgal A. 2017. The Drosophila circuitry of sleep- wake regulation. Current Opinion in Neurobiology
44:243–250. DOI: https://doi.org/10.1016/j.conb.2017.03.004, PMID: 28366532
Aso Y, Grübel K, Busch S, Friedrich AB, Siwanowicz I, Tanimoto H. 2009. The mushroom body of adult
Drosophila characterized by GAL4 drivers. Journal of Neurogenetics 23:156–172. DOI: https://doi.org/10.
1080/01677060802471718, PMID: 19140035
Awasaki T, Lai SL, Ito K, Lee T. 2008. Organization and postembryonic development of glial cells in the adult
central brain of Drosophila. The Journal of Neuroscience 28:13742–13753. DOI: https://doi.org/10.1523/
JNEUROSCI.4844-08.2008, PMID: 19091965
Batista PJ. 2017. The RNA modification N6- methyladenosine and its implications in human disease. Genomics,
Proteomics & Bioinformatics 15:154–163. DOI: https://doi.org/10.1016/j.gpb.2017.03.002, PMID: 28533023
Baumeister AA, Frye GD. 1985. The biochemical basis of the behavioral disorder in the Lesch- Nyhan syndrome.
Neuroscience and Biobehavioral Reviews 9:169–178. DOI: https://doi.org/10.1016/0149-7634(85)90043-0,
PMID: 3925393
Becker JL. 1974a. Purine metabolism in Drosophila melanogaster cells in culture in vitro: purine
interconversion. Biochimie 56:1249–1253. DOI: https://doi.org/10.1016/s0300-9084(74)80018-0, PMID:
4155958
Becker JL. 1974b. Purine metabolism pathways in Drosophila cells grown “in vitro”: phosphoribosyl transferase
activities. Biochimie 56:779–781. DOI: https://doi.org/10.1016/s0300-9084(74)80051-9, PMID: 4451668
Bell S, Kolobova I, Crapper L, Ernst C. 2016. Lesch- nyhan syndrome: Models, theories, and therapies. Molecular
Syndromology 7:302–311. DOI: https://doi.org/10.1159/000449296, PMID: 27920633
Bell S, McCarty V, Peng H, Jefri M, Hettige N, Antonyan L, Crapper L, O’Leary LA, Zhang X, Zhang Y, Wu H,
Sutcliffe D, Kolobova I, Rosenberger TA, Moquin L, Gratton A, Popic J, Gantois I, Stumpf PS, Schuppert AA,
etal. 2021. Lesch- Nyhan disease causes impaired energy metabolism and reduced developmental potential in
midbrain dopaminergic cells. Stem Cell Reports 16:1749–1762. DOI: https://doi.org/10.1016/j.stemcr.2021.06.
003, PMID: 34214487
Bertelli M, Cecchin S, Lapucci C, Jacomelli G, Jinnah HA, Pandolfo M, Micheli V. 2006. Study of the
adenosinergic system in the brain of HPRT knockout mouse (Lesch- Nyhan disease). Clinica Chimica Acta;
International Journal of Clinical Chemistry 373:104–107. DOI: https://doi.org/10.1016/j.cca.2006.05.013, PMID:
16793031
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 29 of 34
Bischof J, Maeda RK, Hediger M, Karch F, Basler K. 2007. An optimized transgenesis system for Drosophila using
germ- line- specific phiC31 integrases. PNAS 104:3312–3317. DOI: https://doi.org/10.1073/pnas.0611511104,
PMID: 17360644
Boison D. 2005. Adenosine and epilepsy: from therapeutic rationale to new therapeutic strategies. The
Neuroscientist 11:25–36. DOI: https://doi.org/10.1177/1073858404269112, PMID: 15632276
Bollée G, Harambat J, Bensman A, Knebelmann B, Daudon M, Ceballos- Picot I. 2012. Adenine
phosphoribosyltransferase deficiency. Clinical Journal of the American Society of Nephrology 7:1521–1527.
DOI: https://doi.org/10.2215/CJN.02320312, PMID: 22700886
Boswell- Casteel RC, Hays FA. 2017. Equilibrative nucleoside transporters: a review. Nucleosides, Nucleotides &
Nucleic Acids 36:7–30. DOI: https://doi.org/10.1080/15257770.2016.1210805, PMID: 27759477
Bou Dib P, Gnägi B, Daly F, Sabado V, Tas D, Glauser DA, Meister P, Nagoshi E. 2014. A conserved role for p48
homologs in protecting dopaminergic neurons from oxidative stress. PLOS Genetics 10:e1004718. DOI:
https://doi.org/10.1371/journal.pgen.1004718, PMID: 25340742
Breese GR, Criswell HE, Duncan GE, Mueller RA. 1990. A dopamine deficiency model of Lesch- Nyhan disease--
the neonatal- 6- OHDA- lesioned rat. Brain Research Bulletin 25:477–484. DOI: https://doi.org/10.1016/0361-
9230(90)90240-z, PMID: 2127238
Brody T, Cravchik A. 2000. Drosophila melanogaster G protein- coupled receptors. The Journal of Cell Biology
150:F83–F88. DOI: https://doi.org/10.1083/jcb.150.2.f83, PMID: 10908591
Cartier P, Hamet M. 1968. Determination of the purine phosphoribosyl transferase activity of human red cells.
Clinica Chimica Acta; International Journal of Clinical Chemistry 20:205–214. DOI: https://doi.org/10.1016/
0009-8981(68)90152-6
Cassar M, Issa AR, Riemensperger T, Petitgas C, Rival T, Coulom H, Iché-Torres M, Han KA, Birman S. 2015. A
dopamine receptor contributes to paraquat- induced neurotoxicity in Drosophila. Human Molecular Genetics
24:197–212. DOI: https://doi.org/10.1093/hmg/ddu430, PMID: 25158689
Ceballos- Picot I, Mockel L, Potier MC, Dauphinot L, Shirley TL, Torero- Ibad R, Fuchs J, Jinnah HA. 2009.
Hypoxanthine- guanine phosphoribosyl transferase regulates early developmental programming of dopamine
neurons: implications for Lesch- Nyhan disease pathogenesis. Human Molecular Genetics 18:2317–2327. DOI:
https://doi.org/10.1093/hmg/ddp164, PMID: 19342420
Ceballos- Picot I, Le Dantec A, Brassier A, Jaïs JP, Ledroit M, Cahu J, Ea HK, Daignan- Fornier B, Pinson B. 2015.
New biomarkers for early diagnosis of Lesch- Nyhan disease revealed by metabolic analysis on a large cohort of
patients. Orphanet Journal of Rare Diseases 10:7. DOI: https://doi.org/10.1186/s13023-014-0219-0, PMID:
25612837
Chang JM, Di Tommaso P, Notredame C. 2014. TCS: a new multiple sequence alignment reliability measure to
estimate alignment accuracy and improve phylogenetic tree reconstruction. Molecular Biology and Evolution
31:1625–1637. DOI: https://doi.org/10.1093/molbev/msu117, PMID: 24694831
Cichewicz K, Garren EJ, Adiele C, Aso Y, Wang Z, Wu M, Birman S, Rubin GM, Hirsh J. 2017. A new brain
dopamine- deficient Drosophila and its pharmacological and genetic rescue. Genes, Brain, and Behavior
16:394–403. DOI: https://doi.org/10.1111/gbb.12353, PMID: 27762066
Cristini S, Navone S, Canzi L, Acerbi F, Ciusani E, Hladnik U, de Gemmis P, Alessandri G, Colombo A, Parati E,
Invernici G. 2010. Human neural stem cells: a model system for the study of Lesch- Nyhan disease neurological
aspects. Human Molecular Genetics 19:1939–1950. DOI: https://doi.org/10.1093/hmg/ddq072, PMID:
20159777
Criswell H, Mueller RA, Breese GR. 1988. Assessment of purine- dopamine interactions in 6- hydroxydopamine-
lesioned rats: evidence for pre- and postsynaptic influences by adenosine. The Journal of Pharmacology and
Experimental Therapeutics 244:493–500 PMID: 3126293.
Darriba D, Taboada GL, Doallo R, Posada D. 2011. ProtTest 3: fast selection of best- fit models of protein
evolution. Bioinformatics 27:1164–1165. DOI: https://doi.org/10.1093/bioinformatics/btr088, PMID: 21335321
Doherty J, Logan MA, Taşdemir OE, Freeman MR. 2009. Ensheathing glia function as phagocytes in the adult
Drosophila brain. The Journal of Neuroscience 29:4768–4781. DOI: https://doi.org/10.1523/JNEUROSCI.
5951-08.2009, PMID: 19369546
Dolezelova E, Nothacker HP, Civelli O, Bryant PJ, Zurovec M. 2007. A Drosophila adenosine receptor activates
cAMP and calcium signaling. Insect Biochemistry and Molecular Biology 37:318–329. DOI: https://doi.org/10.
1016/j.ibmb.2006.12.003, PMID: 17368195
Dunnett SB, Sirinathsinghji DJ, Heavens R, Rogers DC, Kuehn MR. 1989. Monoamine deficiency in a transgenic
(Hprt-) mouse model of Lesch- Nyhan syndrome. Brain Research 501:401–406. DOI: https://doi.org/10.1016/
0006-8993(89)90659-8, PMID: 2573408
Edgar RC. 2004. MUSCLE: a multiple sequence alignment method with reduced time and space complexity.
BMC Bioinformatics 5:113. DOI: https://doi.org/10.1186/1471-2105-5-113, PMID: 15318951
Egami K, Yitta S, Kasim S, Lewers JC, Roberts RC, Lehar M, Jinnah HA. 2007. Basal ganglia dopamine loss due
to defect in purine recycling. Neurobiology of Disease 26:396–407. DOI: https://doi.org/10.1016/j.nbd.2007.
01.010, PMID: 17374562
Engle SJ, Womer DE, Davies PM, Boivin G, Sahota A, Simmonds HA, Stambrook PJ, Tischfield JA. 1996.
HPRT- APRT- deficient mice are not a model for lesch- nyhan syndrome. Human Molecular Genetics 5:1607–1610.
DOI: https://doi.org/10.1093/hmg/5.10.1607, PMID: 8894695
Ernst M, Zametkin AJ, Matochik JA, Pascualvaca D, Jons PH, Hardy K, Hankerson JG, Doudet DJ, Cohen RM.
1996. Presynaptic dopaminergic deficits in Lesch- Nyhan disease. The New England Journal of Medicine
334:1568–1572. DOI: https://doi.org/10.1056/NEJM199606133342403, PMID: 8628337
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 30 of 34
Faville R, Kottler B, Goodhill GJ, Shaw PJ, van Swinderen B. 2015. How deeply does your mutant sleep? probing
arousal to better understand sleep defects in Drosophila. Scientific Reports 5:8454. DOI: https://doi.org/10.
1038/srep08454, PMID: 25677943
Feany MB, Bender WW. 2000. A Drosophila model of parkinson’s disease. Nature 404:394–398. DOI: https://
doi.org/10.1038/35006074, PMID: 10746727
Fergestad T, Bostwick B, Ganetzky B. 2006. Metabolic disruption in Drosophila bang- sensitive seizure mutants.
Genetics 173:1357–1364. DOI: https://doi.org/10.1534/genetics.106.057463, PMID: 16648587
Fernández- Hernández I, Scheenaard E, Pollarolo G, Gonzalez C. 2016. The translational relevance of Drosophila
in drug discovery. EMBO Reports 17:471–472. DOI: https://doi.org/10.15252/embr.201642080, PMID:
26882560
Finger S, Heavens RP, Sirinathsinghji DJ, Kuehn MR, Dunnett SB. 1988. Behavioral and neurochemical evaluation
of a transgenic mouse model of Lesch- Nyhan syndrome. Journal of the Neurological Sciences 86:203–213.
DOI: https://doi.org/10.1016/0022-510x(88)90099-8, PMID: 3221240
Franco R, Ferré S, Agnati L, Torvinen M, Ginés S, Hillion J, Casadó V, Lledó P, Zoli M, Lluis C, Fuxe K. 2000.
Evidence for adenosine/dopamine receptor interactions: indications for heteromerization.
Neuropsychopharmacology 23:S50–S59. DOI: https://doi.org/10.1016/S0893-133X(00)00144-5, PMID:
11008067
Friggi- Grelin F, Coulom H, Meller M, Gomez D, Hirsh J, Birman S. 2003. Targeted gene expression in Drosophila
dopaminergic cells using regulatory sequences from tyrosine hydroxylase. Journal of Neurobiology 54:618–
627. DOI: https://doi.org/10.1002/neu.10185, PMID: 12555273
Fu R, Jinnah HA. 2012. Genotype- phenotype correlations in Lesch- Nyhan disease: moving beyond the gene. The
Journal of Biological Chemistry 287:2997–3008. DOI: https://doi.org/10.1074/jbc.M111.317701, PMID:
22157001
Fu R, Ceballos- Picot I, Torres RJ, Larovere LE, Yamada Y, Nguyen KV, Hegde M, Visser JE, Schretlen DJ,
Nyhan WL, Puig JG, O’Neill PJ, Jinnah HA. 2014. Genotype- phenotype correlations in neurogenetics:
Lesch- Nyhan disease as a model disorder. Brain 137:1282–1303. DOI: https://doi.org/10.1093/brain/awt202,
PMID: 23975452
Fu R, Sutcliffe D, Zhao H, Huang X, Schretlen DJ, Benkovic S, Jinnah HA. 2015. Clinical severity in Lesch- Nyhan
disease: the role of residual enzyme and compensatory pathways. Molecular Genetics and Metabolism
114:55–61. DOI: https://doi.org/10.1016/j.ymgme.2014.11.001, PMID: 25481104
García MG, Puig JG, Torres RJ. 2009. Abnormal adenosine and dopamine receptor expression in lymphocytes of
Lesch- Nyhan patients. Brain, Behavior, and Immunity 23:1125–1131. DOI: https://doi.org/10.1016/j.bbi.2009.
07.006, PMID: 19635551
García MG, Puig JG, Torres RJ. 2012. Adenosine, dopamine and serotonin receptors imbalance in lymphocytes
of Lesch- Nyhan patients. Journal of Inherited Metabolic Disease 35:1129–1135. DOI: https://doi.org/10.1007/
s10545-012-9470-5, PMID: 22403020
Gouy M, Guindon S, Gascuel O. 2010. SeaView version 4: a multiplatform graphical user interface for sequence
alignment and phylogenetic tree building. Molecular Biology and Evolution 27:221–224. DOI: https://doi.org/
10.1093/molbev/msp259, PMID: 19854763
Graham BH, Li Z, Alesii EP, Versteken P, Lee C, Wang J, Craigen WJ. 2010. Neurologic dysfunction and male
infertility in Drosophila porin mutants: a new model for mitochondrial dysfunction and disease. The Journal of
Biological Chemistry 285:11143–11153. DOI: https://doi.org/10.1074/jbc.M109.080317, PMID: 20110367
Gray JH, Owen RP, Giacomini KM. 2004. The concentrative nucleoside transporter family, SLC28. Pflugers Archiv
447:728–734. DOI: https://doi.org/10.1007/s00424-003-1107-y, PMID: 12856181
Guibinga GH, Hrustanovic G, Bouic K, Jinnah HA, Friedmann T. 2012. MicroRNA- mediated dysregulation of
neural developmental genes in HPRT deficiency: clues for Lesch- Nyhan disease? Human Molecular Genetics
21:609–622. DOI: https://doi.org/10.1093/hmg/ddr495, PMID: 22042773
Guindon S, Dufayard JF, Lefort V, Anisimova M, Hordijk W, Gascuel O. 2010. New algorithms and methods to
estimate maximum- likelihood phylogenies: assessing the performance of PhyML 3.0. Systematic Biology
59:307–321. DOI: https://doi.org/10.1093/sysbio/syq010, PMID: 20525638
Harambat J, Bollée G, Daudon M, Ceballos- Picot I, Bensman A. 2012. Adenine phosphoribosyltransferase
deficiency in children. Pediatric Nephrology 27:571–579. DOI: https://doi.org/10.1007/s00467-011-2037-0,
PMID: 22212387
Harkness RA, McCreanor GM, Watts RW. 1988. Lesch- Nyhan syndrome and its pathogenesis: purine
concentrations in plasma and urine with metabolite profiles in CSF. Journal of Inherited Metabolic Disease
11:239–252. DOI: https://doi.org/10.1007/BF01800365, PMID: 3148065
Howlett IC, Rusan ZM, Parker L, Tanouye MA. 2013. Drosophila as a model for intractable epilepsy: gilgamesh
suppresses seizures in para(bss1) heterozygote flies. G3: Genes, Genomes, Genetics 3:1399–1407. DOI:
https://doi.org/10.1534/g3.113.006130, PMID: 23797108
Huber R, Hill SL, Holladay C, Biesiadecki M, Tononi G, Cirelli C. 2004. Sleep homeostasis in Drosophila
melanogaster. Sleep 27:628–639. DOI: https://doi.org/10.1093/sleep/27.4.628, PMID: 15282997
Ipata PL. 2011. Origin, utilization, and recycling of nucleosides in the central nervous system. Advances in
Physiology Education 35:342–346. DOI: https://doi.org/10.1152/advan.00068.2011, PMID: 22139768
Issa AR, Sun J, Petitgas C, Mesquita A, Dulac A, Robin M, Mollereau B, Jenny A, Chérif- Zahar B, Birman S. 2018.
The lysosomal membrane protein LAMP2A promotes autophagic flux and prevents SNCA- induced parkinson
disease- like symptoms in the Drosophila brain. Autophagy 14:1898–1910. DOI: https://doi.org/10.1080/
15548627.2018.1491489, PMID: 29989488
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 31 of 34
Jinnah HA, Gage FH, Friedmann T. 1990. Animal models of Lesch- Nyhan syndrome. Brain Research Bulletin
25:467–475. DOI: https://doi.org/10.1016/0361-9230(90)90239-v, PMID: 2292045
Jinnah HA, Page T, Friedmann T. 1993. Brain purines in a genetic mouse model of Lesch- Nyhan disease. Journal
of Neurochemistry 60:2036–2045. DOI: https://doi.org/10.1111/j.1471-4159.1993.tb03488.x, PMID: 8492116
Jinnah HA, Wojcik BE, Hunt M, Narang N, Lee KY, Goldstein M, Wamsley JK, Langlais PJ, Friedmann T. 1994.
Dopamine deficiency in a genetic mouse model of Lesch- Nyhan disease. The Journal of Neuroscience
14:1164–1175. DOI: https://doi.org/10.1523/JNEUROSCI.14-03-01164.1994, PMID: 7509865
Jinnah HA, Visser JE, Harris JC, Verdu A, Larovere L, Ceballos- Picot I, Gonzalez- Alegre P, Neychev V, Torres RJ,
Dulac O, Desguerre I, Schretlen DJ, Robey KL, Barabas G, Bloem BR, Nyhan W, De Kremer R, Eddey GE,
Puig JG, Reich SG. 2006. Delineation of the motor disorder of Lesch- Nyhan disease. Brain 129:1201–1217.
DOI: https://doi.org/10.1093/brain/awl056, PMID: 16549399
Jinnah HA, Ceballos- Picot I, Torres RJ, Visser JE, Schretlen DJ, Verdu A, Laróvere LE, Chen CJ, Cossu A, Wu CH,
Sampat R, Chang SJ, de Kremer RD, Nyhan W, Harris JC, Reich SG, Puig JG. 2010. Attenuated variants of
Lesch- Nyhan disease. Brain 133:671–689. DOI: https://doi.org/10.1093/brain/awq013, PMID: 20176575
Johnson DH, Friedman TB. 1981. Purine resistant mutants of Drosophila are adenine phosphoribosyltransferase
deficient. Science 212:1035–1036. DOI: https://doi.org/10.1126/science.212.4498.1035, PMID: 17779974
Johnson DH, Friedman TB. 1983. Purine- resistant Drosophila melanogaster result from mutations in the adenine
phosphoribosyltransferase structural gene. PNAS 80:2990–2994. DOI: https://doi.org/10.1073/pnas.80.10.
2990, PMID: 6407004
Johnson DH, Edström JE, Burnett JB, Friedman TB. 1987. Cloning of a Drosophila melanogaster adenine
phosphoribosyltransferase structural gene and deduced amino acid sequence of the enzyme. Gene 59:77–86.
DOI: https://doi.org/10.1016/0378-1119(87)90268-x, PMID: 3125085
Johnson EC, Kazgan N, Bretz CA, Forsberg LJ, Hector CE, Worthen RJ, Onyenwoke R, Brenman JE. 2010.
Altered metabolism and persistent starvation behaviors caused by reduced AMPK function in Drosophila. PLOS
ONE 5:e12799. DOI: https://doi.org/10.1371/journal.pone.0012799, PMID: 20862213
Johnson TA, Jinnah HA, Kamatani N. 2019. Shortage of cellular atp as a cause of diseases and strategies to
enhance atp. Frontiers in Pharmacology 10:98. DOI: https://doi.org/10.3389/fphar.2019.00098, PMID:
30837873
Katoh K, Standley DM. 2013. MAFFT multiple sequence alignment software version 7: improvements in
performance and usability. Molecular Biology and Evolution 30:772–780. DOI: https://doi.org/10.1093/molbev/
mst010, PMID: 23329690
Kelley WN, Rosenbloom FM, Henderson JF, Seegmiller JE. 1967. A specific enzyme defect in gout associated
with overproduction of uric acid. PNAS 57:1735–1739. DOI: https://doi.org/10.1073/pnas.57.6.1735, PMID:
4291947
Kim DS, Palmiter RD. 2008. Interaction of dopamine and adenosine receptor function in behavior: studies with
dopamine- deficient mice. Frontiers in Bioscience 13:2311–2318. DOI: https://doi.org/10.2741/2845, PMID:
17981713
Klarsfeld A, Leloup JC, Rouyer F. 2003. Circadian rhythms of locomotor activity in Drosophila. Behavioural
Processes 64:161–175. DOI: https://doi.org/10.1016/s0376-6357(03)00133-5, PMID: 14556950
Knapp DJ, Breese GR. 2016. The use of perinatal 6- hydroxydopamine to produce a rodent model of lesch- nyhan
disease. Current Topics in Behavioral Neurosciences 29:265–277. DOI: https://doi.org/10.1007/7854_2016_
444, PMID: 27029809
Knight D, Harvey PJ, Iliadi KG, Klose MK, Iliadi N, Dolezelova E, Charlton MP, Zurovec M, Boulianne GL. 2010.
Equilibrative nucleoside transporter 2 regulates associative learning and synaptic function in Drosophila. The
Journal of Neuroscience 30:5047–5057. DOI: https://doi.org/10.1523/JNEUROSCI.6241-09.2010, PMID:
20371825
Kroll JR, Tanouye MA. 2013. Rescue of easily shocked mutant seizure sensitivity in Drosophila adults:
ethanolamine kinase and seizure susceptibility. Journal of Comparative Neurology 521:3500–3507. DOI:
https://doi.org/10.1002/cne.23364
Kroll JR, Wong KG, Siddiqui FM, Tanouye MA. 2015. Disruption of endocytosis with the dynamin mutant
shibirets1 suppresses seizures in Drosophila. Genetics 201:1087–1102. DOI: https://doi.org/10.1534/genetics.
115.177600, PMID: 26341658
Kumar S, Chen D, Sehgal A. 2012. Dopamine acts through cryptochrome to promote acute arousal in
Drosophila. Genes & Development 26:1224–1234. DOI: https://doi.org/10.1101/gad.186338.111, PMID:
22581798
Lahaye C, Augé F, Soubrier M, Ceballos- Picot I. 2014. New mutation affecting hypoxanthine
phosphoribosyltransferase responsible for severe tophaceous gout. The Journal of Rheumatology 41:1252–
1254. DOI: https://doi.org/10.3899/jrheum.131168, PMID: 24882866
Lartillot N, Lepage T, Blanquart S. 2009. PhyloBayes 3: a bayesian software package for phylogenetic
reconstruction and molecular dating. Bioinformatics 25:2286–2288. DOI: https://doi.org/10.1093/
bioinformatics/btp368, PMID: 19535536
Lee G, Kikuno K, Bahn JH, Kim KM, Park JH. 2013. Dopamine D2 receptor as a cellular component controlling
nocturnal hyperactivities in Drosophila melanogaster. Chronobiology International 30:443–459. DOI: https://
doi.org/10.3109/07420528.2012.741169, PMID: 23286280
Lesch M, Nyhan WL. 1964. A familial disorder of uric acid metabolism and central nervous system function. The
American Journal of Medicine 36:561–570. DOI: https://doi.org/10.1016/0002-9343(64)90104-4, PMID:
14142409
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 32 of 34
Liu C, Plaçais PY, Yamagata N, Pfeiffer BD, Aso Y, Friedrich AB, Siwanowicz I, Rubin GM, Preat T, Tanimoto H.
2012. A subset of dopamine neurons signals reward for odour memory in Drosophila. Nature 488:512–516.
DOI: https://doi.org/10.1038/nature11304, PMID: 22810589
Lloyd KG, Hornykiewicz O, Davidson L, Shannak K, Farley I, Goldstein M, Shibuya M, Kelley WN, Fox IH. 1981.
Biochemical evidence of dysfunction of brain neurotransmitters in the Lesch- Nyhan syndrome. The New
England Journal of Medicine 305:1106–1111. DOI: https://doi.org/10.1056/NEJM198111053051902, PMID:
6117011
López JM. 2008. Is ZMP the toxic metabolite in Lesch- Nyhan disease? Medical Hypotheses 71:657–663. DOI:
https://doi.org/10.1016/j.mehy.2008.06.033, PMID: 18710792
López JM, Outtrim EL, Fu R, Sutcliffe DJ, Torres RJ, Jinnah HA. 2020. Physiological levels of folic acid reveal
purine alterations in Lesch- Nyhan disease. PNAS 117:12071–12079. DOI: https://doi.org/10.1073/pnas.
2003475117, PMID: 32430324
Machado J, Abdulla P, Hanna WJB, Hilliker AJ, Coe IR. 2007. Genomic analysis of nucleoside transporters in
diptera and functional characterization of DmENT2, a Drosophila equilibrative nucleoside transporter.
Physiological Genomics 28:337–347. DOI: https://doi.org/10.1152/physiolgenomics.00087.2006, PMID:
17090699
Madeo A, Di Rocco M, Brassier A, Bahi- Buisson N, De Lonlay P, Ceballos- Picot I. 2019. Clinical, biochemical
and genetic characteristics of a cohort of 101 French and Italian patients with HPRT deficiency. Molecular
Genetics and Metabolism 127:147–157. DOI: https://doi.org/10.1016/j.ymgme.2019.06.001, PMID:
31182398
Mao Z, Davis RL. 2009. Eight different types of dopaminergic neurons innervate the Drosophila mushroom body
neuropil: anatomical and physiological heterogeneity. Frontiers in Neural Circuits 3:5. DOI: https://doi.org/10.
3389/neuro.04.005.2009, PMID: 19597562
Marks JF, Baum J, Keele DK, Kay JL, MacFarlen A. 1968. Lesch- Nyhan syndrome treated from the early neonatal
period. Pediatrics 42:357–359 PMID: 5663742.
Masino SA, Kawamura M, Ruskin DN. 2014. Adenosine receptors and epilepsy: current evidence and future
potential. International Review of Neurobiology 119:233–255. DOI: https://doi.org/10.1016/B978-0-12-
801022-8.00011-8, PMID: 25175969
Mazaud D, Kottler B, Gonçalves- Pimentel C, Proelss S, Tüchler N, Deneubourg C, Yuasa Y, Diebold C,
Jungbluth H, Lai EC, Hirth F, Giangrande A, Fanto M. 2019. Transcriptional regulation of the glutamate/gaba/
glutamine cycle in adult glia controls motor activity and seizures in Drosophila. The Journal of Neuroscience
39:5269–5283. DOI: https://doi.org/10.1523/JNEUROSCI.1833-18.2019, PMID: 31064860
McGuire SE, Le PT, Osborn AJ, Matsumoto K, Davis RL. 2003. Spatiotemporal rescue of memory dysfunction in
Drosophila. Science 302:1765–1768. DOI: https://doi.org/10.1126/science.1089035, PMID: 14657498
Meek S, Thomson AJ, Sutherland L, Sharp MGF, Thomson J, Bishop V, Meddle SL, Gloaguen Y, Weidt S,
Singh- Dolt K, Buehr M, Brown HK, Gill AC, Burdon T. 2016. Reduced levels of dopamine and altered
metabolism in brains of HPRT knock- out rats: a new rodent model of Lesch- Nyhan Disease. Scientific Reports
6:25592. DOI: https://doi.org/10.1038/srep25592, PMID: 27185277
Miller S, Collins JM. 1973. Metabolic purine pathways in the developing ovary of the housefly, musca domestica.
Comparative Biochemistry and Physiology Part B 44:1153–1163. DOI: https://doi.org/10.1016/0305-0491(73)
90267-8
Miñana MD, Portolés M, Jordá A, Grisolía S. 1984. Lesch- Nyhan syndrome, caffeine model: increase of purine
and pyrimidine enzymes in rat brain. Journal of Neurochemistry 43:1556–1560. DOI: https://doi.org/10.1111/j.
1471-4159.1984.tb06078.x, PMID: 6149265
Mizuno T, Ohta R, Kodama K, Kitazumi E, Minejima N, Takeishi M, Segawa M. 1979. Self- mutilation and sleep
stage in the Lesch- Nyhan syndrome. Brain & Development 1:121–125 PMID: 233225.
Nall AH, Shakhmantsir I, Cichewicz K, Birman S, Hirsh J, Sehgal A. 2016. Caffeine promotes wakefulness via
dopamine signaling in Drosophila. Scientific Reports 6:20938. DOI: https://doi.org/10.1038/srep20938, PMID:
26868675
Nyhan WL. 1997. The recognition of Lesch- Nyhan syndrome as an inborn error of purine metabolism. Journal of
Inherited Metabolic Disease 20:171–178. DOI: https://doi.org/10.1023/a:1005348504512, PMID: 9211189
Nyhan WL. 2000. Dopamine function in Lesch- Nyhan disease. Environmental Health Perspectives 108 Suppl
3:409–411. DOI: https://doi.org/10.1289/ehp.00108s3409, PMID: 10852837
Nyhan WL. 2014. Nucleotide synthesis via salvage pathway. Nyhan WL (Ed). Encyclopedia of Life Sciences. Wiley.
p. 1–10. DOI: https://doi.org/10.1002/047001590X
Oriel C, Lasko P. 2018. Recent developments in using Drosophila as a model for human genetic disease.
International Journal of Molecular Sciences 19:2041. DOI: https://doi.org/10.3390/ijms19072041, PMID:
30011838
Ozeir M, Huyet J, Burgevin MC, Pinson B, Chesney F, Remy JM, Siddiqi AR, Lupoli R, Pinon G, Saint- Marc C,
Gibert JF, Morales R, Ceballos- Picot I, Barouki R, Daignan- Fornier B, Olivier- Bandini A, Augé F, Nioche P. 2019.
Structural basis for substrate selectivity and nucleophilic substitution mechanisms in human adenine
phosphoribosyltransferase catalyzed reaction. The Journal of Biological Chemistry 294:11980–11991. DOI:
https://doi.org/10.1074/jbc.RA119.009087, PMID: 31160323
Papanikolopoulou K, Mudher A, Skoulakis E. 2019. An assessment of the translational relevance of Drosophila in
drug discovery. Expert Opinion on Drug Discovery 14:303–313. DOI: https://doi.org/10.1080/17460441.2019.
1569624, PMID: 30664368
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 33 of 34
Parker L, Howlett IC, Rusan ZM, Tanouye MA. 2011. Seizure and epilepsy: studies of seizure disorders in
Drosophila. International Review of Neurobiology 99:1–21. DOI: https://doi.org/10.1016/B978-0-12-387003-2.
00001-X, PMID: 21906534
Pastor- Anglada M, Pérez- Torras S. 2018. Emerging roles of nucleoside transporters. Frontiers in Pharmacology
9:606. DOI: https://doi.org/10.3389/fphar.2018.00606, PMID: 29928232
Pech U, Pooryasin A, Birman S, Fiala A. 2013. Localization of the contacts between kenyon cells and aminergic
neurons in the Drosophila melanogaster brain using SplitGFP reconstitution. The Journal of Comparative
Neurology 521:3992–4026. DOI: https://doi.org/10.1002/cne.23388, PMID: 23784863
Perrimon N, Bonini NM, Dhillon P. 2016. Fruit flies on the front line: the translational impact of Drosophila.
Disease Models & Mechanisms 9:229–231. DOI: https://doi.org/10.1242/dmm.024810, PMID: 26935101
Plaçais PY, de Tredern É, Scheunemann L, Trannoy S, Goguel V, Han KA, Isabel G, Preat T. 2017. Upregulated
energy metabolism in the Drosophila mushroom body is the trigger for long- term memory. Nature
Communications 8:15510. DOI: https://doi.org/10.1038/ncomms15510, PMID: 28580949
Pütz SM, Kram J, Rauh E, Kaiser S, Toews R, Lueningschroer- Wang Y, Rieger D, Raabe T. 2021. Loss of p21-
activated kinase Mbt/PAK4 causes parkinson- like phenotypes in Drosophila. Disease Models & Mechanisms
14:dmm047811. DOI: https://doi.org/10.1242/dmm.047811, PMID: 34125184
Rahmani Z, Surabhi S, Rojo- Cortés F, Dulac A, Jenny A, Birman S. 2022. Lamp1 deficiency enhances sensitivity to
α-synuclein and oxidative stress in Drosophila models of parkinson disease. International Journal of Molecular
Sciences 23:13078. DOI: https://doi.org/10.3390/ijms232113078, PMID: 36361864
Riemensperger T, Isabel G, Coulom H, Neuser K, Seugnet L, Kume K, Iché-Torres M, Cassar M, Strauss R,
Preat T, Hirsh J, Birman S. 2011. Behavioral consequences of dopamine deficiency in the Drosophila central
nervous system. PNAS 108:834–839. DOI: https://doi.org/10.1073/pnas.1010930108, PMID: 21187381
Riemensperger T, Issa AR, Pech U, Coulom H, Nguyn MV, Cassar M, Jacquet M, Fiala A, Birman S. 2013. A
single dopamine pathway underlies progressive locomotor deficits in a Drosophila model of parkinson disease.
Cell Reports 5:952–960. DOI: https://doi.org/10.1016/j.celrep.2013.10.032, PMID: 24239353
Rival T, Soustelle L, Strambi C, Besson MT, Iché M, Birman S. 2004. Decreasing glutamate buffering capacity
triggers oxidative stress and neuropil degeneration in the Drosophila brain. Current Biology 14:599–605. DOI:
https://doi.org/10.1016/j.cub.2004.03.039, PMID: 15062101
Ruillier V, Tournois J, Boissart C, Lasbareilles M, Mahé G, Chatrousse L, Cailleret M, Peschanski M, Benchoua A.
2020. Rescuing compounds for Lesch- Nyhan disease identified using stem cell- based phenotypic screening.
JCI Insight 5:e132094. DOI: https://doi.org/10.1172/jci.insight.132094, PMID: 31990683
Saito Y, Takashima S. 2000. Neurotransmitter changes in the pathophysiology of Lesch- Nyhan syndrome. Brain &
Development 22 Suppl 1:S122–S131. DOI: https://doi.org/10.1016/s0387-7604(00)00143-1, PMID: 10984673
Sankar N, Machado J, Abdulla P, Hilliker AJ, Coe IR. 2002. Comparative genomic analysis of equilibrative
nucleoside transporters suggests conserved protein structure despite limited sequence identity. Nucleic Acids
Research 30:4339–4350. DOI: https://doi.org/10.1093/nar/gkf564, PMID: 12384580
Saras A, Tanouye MA. 2016. Seizure suppression by high temperature via camp modulation in Drosophila. G3:
Genes, Genomes, Genetics 6:3381–3387. DOI: https://doi.org/10.1534/g3.116.034629, PMID: 27558668
Sass JK, Itabashi HH, Dexter RA. 1965. Juvenile gout with brain involvement. Archives of Neurology 13:639–655.
DOI: https://doi.org/10.1001/archneur.1965.00470060075008, PMID: 5850672
Schram KH. 1998. Urinary nucleosides. Mass Spectrometry Reviews 17:131–251. DOI: https://doi.org/10.1002/(
SICI)1098-2787(1998)17:3<131::AID-MAS1>3.0.CO;2-O, PMID: 10095827
Schretlen DJ, Ward J, Meyer SM, Yun J, Puig JG, Nyhan WL, Jinnah HA, Harris JC. 2005. Behavioral aspects of
Lesch- Nyhan disease and its variants. Developmental Medicine and Child Neurology 47:673–677. DOI: https://
doi.org/10.1017/S0012162205001374, PMID: 16174310
Seegmiller JE, Rosenbloom FM, Kelley WN. 1967. Enzyme defect associated with a sex- linked human
neurological disorder and excessive purine synthesis. Science 155:1682–1684. DOI: https://doi.org/10.1126/
science.155.3770.1682, PMID: 6020292
Sinakevitch I, Grau Y, Strausfeld NJ, Birman S. 2010. Dynamics of glutamatergic signaling in the mushroom body
of young adult Drosophila. Neural Development 5:10. DOI: https://doi.org/10.1186/1749-8104-5-10, PMID:
20370889
Sinha SC, Smith JL. 2001. The PRT protein family. Current Opinion in Structural Biology 11:733–739. DOI:
https://doi.org/10.1016/s0959-440x(01)00274-3, PMID: 11751055
Sinnett SE, Brenman JE. 2016. The role of AMPK in Drosophila melanogaster. Experientia. Supplementum
107:389–401. DOI: https://doi.org/10.1007/978-3-319-43589-3_16, PMID: 27812989
Smith DW, Friedmann T. 2000. Characterization of the dopamine defect in primary cultures of dopaminergic
neurons from hypoxanthine phosphoribosyltransferase knockout mice. Molecular Therapy 1:486–491. DOI:
https://doi.org/10.1006/mthe.2000.0057, PMID: 10933970
Soliman AM, Fathalla AM, Moustafa AA. 2018. Adenosine role in brain functions: pathophysiological influence
on parkinson’s disease and other brain disorders. Pharmacological Reports 70:661–667. DOI: https://doi.org/
10.1016/j.pharep.2018.02.003, PMID: 29909246
Song J, Tanouye MA. 2008. From bench to drug: human seizure modeling using Drosophila. Progress in
Neurobiology 84:182–191. DOI: https://doi.org/10.1016/j.pneurobio.2007.10.006, PMID: 18063465
Stenesen D, Suh JM, Seo J, Yu K, Lee KS, Kim JS, Min KJ, Graff JM. 2013. Adenosine nucleotide biosynthesis
and AMPK regulate adult life span and mediate the longevity benefit of caloric restriction in flies. Cell
Metabolism 17:101–112. DOI: https://doi.org/10.1016/j.cmet.2012.12.006, PMID: 23312286
Research article Genetics and Genomics | Neuroscience
Petitgas etal. eLife 2023;12:RP88510. DOI: https://doi.org/10.7554/eLife.88510 34 of 34
Sun J, Xu AQ, Giraud J, Poppinga H, Riemensperger T, Fiala A, Birman S. 2018. Neural control of startle- induced
locomotion by the mushroom bodies and associated neurons in Drosophila. Frontiers in Systems Neuroscience
12:6. DOI: https://doi.org/10.3389/fnsys.2018.00006, PMID: 29643770
Sutcliffe DJ, Dinasarapu AR, Visser JE, den Hoed J, Seifar F, Joshi P, Ceballos- Picot I, Sardar T, Hess EJ, Sun YV,
Wen Z, Zwick ME, Jinnah HA. 2021. Induced pluripotent stem cells from subjects with Lesch- Nyhan disease.
Scientific Reports 11:8523. DOI: https://doi.org/10.1038/s41598-021-87955-9, PMID: 33875724
Tas D, Stickley L, Miozzo F, Koch R, Loncle N, Sabado V, Gnägi B, Nagoshi E. 2018. Parallel roles of transcription
factors dFOXO and FER2 in the development and maintenance of dopaminergic neurons. PLOS Genetics
14:e1007271. DOI: https://doi.org/10.1371/journal.pgen.1007271, PMID: 29529025
Torres RJ, Deantonio I, Prior C, Puig JG. 2004. Adenosine transport in peripheral blood lymphocytes from
Lesch- Nyhan patients. The Biochemical Journal 377:733–739. DOI: https://doi.org/10.1042/BJ20031035,
PMID: 14572307
Torres RJ, Prior C, Puig JG. 2007a. Efficacy and safety of allopurinol in patients with hypoxanthine- guanine
phosphoribosyltransferase deficiency. Metabolism 56:1179–1186. DOI: https://doi.org/10.1016/j.metabol.
2007.04.013, PMID: 17697859
Torres RJ, Puig JG. 2007b. Hypoxanthine- guanine phosophoribosyltransferase (HPRT) deficiency: Lesch- Nyhan
syndrome. Orphanet Journal of Rare Diseases 2:48. DOI: https://doi.org/10.1186/1750-1172-2-48, PMID:
18067674
Trifinopoulos J, Nguyen LT, von Haeseler A, Minh BQ. 2016. W- IQ- TREE: a fast online phylogenetic tool for
maximum likelihood analysis. Nucleic Acids Research 44:W232–W235. DOI: https://doi.org/10.1093/nar/
gkw256, PMID: 27084950
Vaccaro A, Issa AR, Seugnet L, Birman S, Klarsfeld A. 2017. Drosophila clock is required in brain pacemaker
neurons to prevent premature locomotor aging independently of its circadian function. PLOS Genetics
13:e1006507. DOI: https://doi.org/10.1371/journal.pgen.1006507, PMID: 28072817
Visser JE, Bär PR, Jinnah HA. 2000. Lesch- Nyhan disease and the basal ganglia. Brain Research. Brain Research
Reviews 32:449–475. DOI: https://doi.org/10.1016/s0165-0173(99)00094-6, PMID: 10760551
Waddell S. 2010. Dopamine reveals neural circuit mechanisms of fly memory. Trends in Neurosciences 33:457–
464. DOI: https://doi.org/10.1016/j.tins.2010.07.001, PMID: 20701984
Wallrath LL, Burnett JB, Friedman TB. 1990. Molecular characterization of the Drosophila melanogaster urate
oxidase gene, an ecdysone- repressible gene expressed only in the malpighian tubules. Molecular and Cellular
Biology 10:5114–5127. DOI: https://doi.org/10.1128/mcb.10.10.5114-5127.1990, PMID: 2118989
Weltha L, Reemmer J, Boison D. 2019. The role of adenosine in epilepsy. Brain Research Bulletin 151:46–54.
DOI: https://doi.org/10.1016/j.brainresbull.2018.11.008, PMID: 30468847
Witteveen JS, Loopstok SR, Ballesteros LL, Boonstra A, van Bakel NHM, van Boekel WHP, Martens GJM,
Visser JE, Kolk SM. 2022. HGprt deficiency disrupts dopaminergic circuit development in a genetic mouse
model of Lesch- Nyhan disease. Cellular and Molecular Life Sciences 79:341. DOI: https://doi.org/10.1007/
s00018-022-04326-x, PMID: 35660973
Wong DF, Harris JC, Naidu S, Yokoi F, Marenco S, Dannals RF, Ravert HT, Yaster M, Evans A, Rousset O,
Bryan RN, Gjedde A, Kuhar MJ, Breese GR. 1996. Dopamine transporters are markedly reduced in Lesch-
Nyhan disease in vivo. PNAS 93:5539–5543. DOI: https://doi.org/10.1073/pnas.93.11.5539, PMID: 8643611
Wu MN, Ho K, Crocker A, Yue Z, Koh K, Sehgal A. 2009. The effects of caffeine on sleep in Drosophila require
PKA activity, but not the adenosine receptor. The Journal of Neuroscience 29:11029–11037. DOI: https://doi.
org/10.1523/JNEUROSCI.1653-09.2009, PMID: 19726661
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Parkinson disease (PD) is a common neurodegenerative condition affecting people predominantly at old age that is characterized by a progressive loss of midbrain dopaminergic neurons and by the accumulation of α-synuclein-containing intraneuronal inclusions known as Lewy bodies. Defects in cellular degradation processes such as the autophagy-lysosomal pathway are suspected to be involved in PD progression. The mammalian Lysosomal-associated membrane proteins LAMP1 and LAMP2 are transmembrane glycoproteins localized in lysosomes and late endosomes that are involved in autophagosome/lysosome maturation and function. Here, we show that the lack of Drosophila Lamp1, the homolog of LAMP1 and LAMP2, severely increased fly susceptibility to paraquat, a pro-oxidant compound known as a potential PD inducer in humans. Moreover, the loss of Lamp1 also exacerbated the progressive locomotor defects induced by the expression of PD-associated mutant α-synuclein A30P (α-synA30P) in dopaminergic neurons. Remarkably, the ubiquitous re-expression of Lamp1 in a mutant context fully suppressed all these defects and conferred significant resistance towards both PD factors above that of wild-type flies. Immunostaining analysis showed that the brain levels of α-synA30P were unexpectedly decreased in young adult Lamp1-deficient flies expressing this protein in comparison to non-mutant controls. This suggests that Lamp1 could neutralize α-synuclein toxicity by promoting the formation of non-pathogenic aggregates in neurons. Overall, our findings reveal a novel role for Drosophila Lamp1 in protecting against oxidative stress and α-synuclein neurotoxicity in PD models, thus furthering our understanding of the function of its mammalian homologs.
Article
Full-text available
In Lesch–Nyhan disease (LND), deficiency of the purine salvage enzyme hypoxanthine guanine phosphoribosyl transferase (HGprt) leads to a characteristic neurobehavioral phenotype dominated by dystonia, cognitive deficits and incapacitating self-injurious behavior. It has been known for decades that LND is associated with dysfunction of midbrain dopamine neurons, without overt structural brain abnormalities. Emerging post mortem and in vitro evidence supports the hypothesis that the dopaminergic dysfunction in LND is of developmental origin, but specific pathogenic mechanisms have not been revealed. In the current study, HGprt deficiency causes specific neurodevelopmental abnormalities in mice during embryogenesis, particularly affecting proliferation and migration of developing midbrain dopamine (mDA) neurons. In mutant embryos at E14.5, proliferation was increased, accompanied by a decrease in cell cycle exit and the distribution and orientation of dividing cells suggested a premature deviation from their migratory route. An abnormally structured radial glia-like scaffold supporting this mDA neuronal migration might lie at the basis of these abnormalities. Consequently, these abnormalities were associated with an increase in area occupied by TH ⁺ cells and an abnormal mDA subpopulation organization at E18.5. Finally, dopaminergic innervation was disorganized in prefrontal and decreased in HGprt deficient primary motor and somatosensory cortices. These data provide direct in vivo evidence for a neurodevelopmental nature of the brain disorder in LND. Future studies should not only focus the specific molecular mechanisms underlying the reported neurodevelopmental abnormalities, but also on optimal timing of therapeutic interventions to rescue the DA neuron defects, which may also be relevant for other neurodevelopmental disorders.
Article
Full-text available
Mutations in HPRT1, a gene encoding a rate-limiting enzyme for purine salvage, cause Lesch-Nyhan disease which is characterized by self-injury and motor impairments. We leveraged stem cell and genetic engineering technologies to model the disease in isogenic and patient-derived forebrain and midbrain cell types. Dopaminergic progenitor cells deficient in HPRT showed decreased intensity of all developmental cell-fate markers measured. Metabolic analyses revealed significant loss of all purine derivatives, except hypoxanthine, and impaired glycolysis and oxidative phosphorylation. real-time glucose tracing demonstrated increased shunting to the pentose phosphate pathway for de novo purine synthesis at the expense of ATP production. Purine depletion in dopaminergic progenitor cells resulted in loss of RHEB, impairing mTORC1 activation. These data demonstrate dopaminergic-specific effects of purine salvage deficiency and unexpectedly reveal that dopaminergic progenitor cells are programmed to a high-energy state prior to higher energy demands of terminally differentiated cells.
Article
Full-text available
Parkinson's disease (PD) provokes bradykinesia, resting tremor, rigidity and postural instability but also non-motor symptoms such as depression, anxiety, sleep and cognitive impairments. Similar phenotypes can be induced in Drosophila melanogaster through modification of PD-relevant genes or the administration of PD-inducing toxins. Recent studies correlated deregulation of human p21-activated kinase PAK4 with PD, leaving open the question of a causative relationship of mutations in this gene for manifestation of PD symptoms. To figure out whether flies lacking the PAK4 homolog Mushroom bodies tiny (Mbt) show PD-like phenotypes, we tested for a variety of PD criteria. Here we demonstrate that mbt mutant flies show PD-like phenotypes including age-dependent movement deficits, reduced life expectancy and fragmented sleep. They also react to a stressful situation with higher immobility, indicating an influence of Mbt on emotional behavior. Loss of Mbt function has a negative effect on the number of dopaminergic protocerebral anterior medial (PAM) neurons, most likely caused by a proliferation defect of neural progenitors. The age-dependent movement deficits are not accompanied by a corresponding further loss of PAM neurons. Previous studies highlighted the importance of a small PAM subgroup for age-dependent PD motor impairments. We show that impaired motor skills are caused by lack of Mbt in this PAM subgroup. In addition, a broader re-expression of Mbt in PAM neurons improves life expectancy. Conversely, selective Mbt knockout in the same cells shortens life span. We conclude that mutations in Mbt/PAK4 can play a causative role in the development of Parkinson's disease phenotypes.
Article
Full-text available
Lesch-Nyhan disease (LND) is an inherited disorder caused by pathogenic variants in the HPRT1 gene, which encodes the purine recycling enzyme hypoxanthine–guanine phosphoribosyltransferase (HGprt). We generated 6 induced pluripotent stem cell (iPSC) lines from 3 individuals with LND, along with 6 control lines from 3 normal individuals. All 12 lines had the characteristics of pluripotent stem cells, as assessed by immunostaining for pluripotency markers, expression of pluripotency genes, and differentiation into the 3 primary germ cell layers. Gene expression profiling with RNAseq demonstrated significant heterogeneity among the lines. Despite this heterogeneity, several anticipated abnormalities were readily detectable across all LND lines, including reduced HPRT1 mRNA. Several unexpected abnormalities were also consistently detectable across the LND lines, including decreases in FAR2P1 and increases in RNF39. Shotgun proteomics also demonstrated several expected abnormalities in the LND lines, such as absence of HGprt protein. The proteomics study also revealed several unexpected abnormalities across the LND lines, including increases in GNAO1 decreases in NSE4A. There was a good but partial correlation between abnormalities revealed by the RNAseq and proteomics methods. Finally, functional studies demonstrated LND lines had no HGprt enzyme activity and resistance to the toxic pro-drug 6-thioguanine. Intracellular purines in the LND lines were normal, but they did not recycle hypoxanthine. These cells provide a novel resource to reveal insights into the relevance of heterogeneity among iPSC lines and applications for modeling LND.
Article
Full-text available
The reversible adenine phosphoribosyltransferase enzyme (APRT) is essential for purine homeostasis in prokaryotes and eukaryotes. In humans, APRT (hAPRT) is the only enzyme known to produce AMP in cells from dietary adenine. APRT can also process adenine analogs, which are involved in plant development or neuronal homeostasis. However, the molecular mechanism underlying substrate specificity of APRT and catalysis in both directions of the reaction remains poorly understood. Here we present the crystal structures of hAPRT complexed to three cellular nucleotide analogs (Hypoxanthine, IMP and GMP) that we compare to the phosphate-bound enzyme. We established that binding to hAPRT is substrate-shape-specific in the forward reaction, while it is base-specific in the reverse reaction. Furthermore, a quantum mechanics/molecular mechanics (QM/MM) analysis suggests that the forward reaction is mainly a nucleophilic substitution of type 2 (SN2) with a mix of SN1-type molecular mechanism. Based on our structural analysis, a magnesium-assisted SN2-type mechanism would be involved in the reverse reaction. These results provide a framework for understanding the molecular mechanism and substrate discrimination in both directions by APRTs. This knowledge can play an instrumental role in the design of inhibitors, such as antiparasitic agents, or adenine-based substrates.
Article
The urate oxidase (UO) gene of Drosophila melanogaster is expressed during the third-instar larval and adult stages, exclusively within a subset of cells of the Malpighian tubules. The UO gene contains a 69-base-pair intron and encodes mature mRNAs of 1,224, 1,227, and 1,244 nucleotides, depending on the site of 3' endonucleolytic cleavage prior to polyadenylation. A direct repeat, 5'-AAGTGAGAGTGAT-3', is the proposed cis-regulatory element involved in 20-hydroxyecdysone repression of the UO gene. The deduced amino acid sequences of UO of D. melanogaster, rat, mouse, and pig and uricase II of soybean show 32 to 38% identity, with 22% of amino acid residues identical in all species. With use of P-element-mediated germ line transformation, 826 base pairs 5' and approximately 1,200 base pairs 3' of the D. melanogaster UO transcribed region contain all of the cis elements allowing for appropriate temporal regulation and Malpighian tubule-specific expression of the UO gene.
Article
Although adenosine plays a key role in multiple motor, affective, and cognitive processes, it has received less attention in the neuroscience field compared to other neurotransmitters (e.g., dopamine). In this review, we highlight the role of adenosine in behavior as well as its interaction with other neurotransmitters, such as dopamine. We also discuss brain disorders impacted by alterations to adenosine, and how targeting adenosine can ameliorate Parkinson’s disease motor symptoms. We also discuss the role of caffeine (as an adenosine antagonist) on cognition as well as a neuroprotective agent against Parkinson’s disease (PD).
Article
Significance Fibroblasts from patients with Lesch-Nyhan disease (LND) do not show any significant nucleotide alteration when they are maintained in regular culture medium, which typically contains artificially high levels of folic acid. However, they show ATP depletion and accumulation of 5-aminoimidazole-4-carboxamide riboside 5′-monophosphate (ZMP; an intermediary of the novo purine biosynthetic pathway) when the culture medium has physiological levels of folic acid. A derivative of ZMP (AICAr) is present in the urine and the cerebrospinal fluid of patients with LND, but not in control individuals, suggesting that a similar situation may occur in vivo. The manipulation of folic acid levels may therefore provide a valuable strategy to further study the pathogenesis of LND.
Article
Lesch-Nyhan disease (LND) is a rare monogenic disease caused by deficiency of the salvage pathway enzyme hypoxanthine-guanine phosphoribosyltransferase (HGPRT) and is characterized by severe neuropsychiatric symptoms that currently cannot be treated. Predictive in vivo models are lacking for screening and evaluating candidate drugs because LND-associated neurological symptoms are not recapitulated in HGPRT-deficient animals. Here, we used human neural stem cells and neurons derived from induced pluripotent stem cells (iPSCs) of children affected by LND to identify neural phenotypes of interest associated with HGPRT deficiency to develop a target-agnostic-based drug screening system. We screened more than 3000 molecules and identified 6 pharmacological compounds, all possessing an adenosine moiety, that corrected HGPRT deficiency-associated neuronal phenotypes by promoting metabolism compensations in an HGPRT-independent manner. This included S-adenosylmethionine (SAM), a compound that had already been used as a compassionate approach to ease the neuropsychiatric symptoms in LND. Interestingly, these compounds compensate abnormal metabolism in a manner complementary to the gold standard allopurinol and can be provided to LND patients via simple food supplementation. This experimental paradigm can be easily adapted to other metabolic disorders affecting normal brain development and functioning in the absence of a relevant animal model.