Skip to main content
PLOS Biology logoLink to PLOS Biology
. 2025 Aug 13;23(8):e3003338. doi: 10.1371/journal.pbio.3003338

Mitochondrial ROS and HIF-1α signaling mediate synaptic plasticity in the critical period

Daniel Sobrido-Cameán 1,*, Bramwell Coulson 2, Michael Miller 1, Matthew C W Oswald 1, Tom Pettini 1, David M D Bailey 1, Richard A Baines 2,*, Matthias Landgraf 1,*
Editor: Timothy Mosca3
PMCID: PMC12367176  PMID: 40802851

Abstract

As developing networks transition from spontaneous irregular to patterned activity, they undergo plastic tuning phases, termed “critical periods”; “critical” because disturbances during these phases can lead to lasting changes in network development and output. Critical periods are common to developing nervous systems, with analogous features shared from insects to mammals, yet the core signaling mechanisms that underlie cellular critical period plasticity have remained elusive. To identify these, we exploited the Drosophila larval locomotor network as an advantageous model system. It has a defined critical period and offers unparalleled access to identified network elements, including the neuromuscular junction as a model synapse. We find that manipulations of a single motoneuron or muscle cell during the critical period lead to predictable, and permanent, cell-specific changes. This demonstrates that critical period adjustments occur at a single-cell level. Mechanistically, we identified mitochondrial reactive oxygen species (ROS) as causative. Specifically, we show that ROS produced by Complex-I of the mitochondrial electron transport chain, generated by the reverse flow of electrons, is necessary and instructive for critical period-regulated plasticity. Downstream of ROS, we identified the Drosophila homologue of hypoxia-inducible factor (HIF-1α), as required for transducing the mitochondrial ROS signal to the nucleus. This signaling axis is also sufficient to cell autonomously specify changes in neuronal properties and animal behavior but, again, only when activated during the embryonic critical period. Thus, we have identified specific mitochondrial ROS and HIF-1α as primary signals that mediate critical period plasticity.


Developing motor networks rely on brief windows of plasticity to fine-tune function. This study identifies mitochondrial reactive oxygen species and HIF-1α as instructive signals that drive lasting neural and behavioral changes during these critical periods, linking metabolism to circuit development.

Introduction

The emergence of network function is arguably among the least well-understood aspects of nervous system development. It is well known that the emergence of network function, which occurs during late stages of nervous system development, correlates with a phase of adjustment during which networks transition from seemingly irregular to more patterned activity. These phases have been termed ‘sensitive’ or ‘critical periods’ and are commonly viewed as periods of heightened plasticity. Importantly, they are also highly sensitive to perturbations: errors induced during the critical period generally have lasting impact, with subsequent plasticity mechanisms unable to correct them [13]. Thus, “decisions” made during the critical period can specify neuronal properties over the long-term, and abnormal critical period experiences likely contribute to epilepsy and other neurodevelopmental conditions such as schizophrenia and autism spectrum disorder [4,5]. Though critical periods have been studied for decades, notably in mammalian sensory systems [3,6,7], many questions remain. Here, we focused on two fundamental unresolved questions: whether critical periods are principally a network phenomenon, or a property of individual cells and, secondly, the identity of the primary signals directing critical period plasticity.

A difficulty in understanding critical periods has been the complexity of mammalian circuitry, including the realization that different cell types can behave differently [6]. Therefore, we exploit a much simpler experimental system, the fruit fly, Drosophila melanogaster, which has enabled the identification of numerous genes and mechanisms that orchestrate the development and function of the nervous system [810]. Importantly, the larval locomotor network has a well-defined critical period in late embryogenesis, with features that suggest it is analogous to mammalian critical periods [11,12]. As is the case for mammals, Drosophila critical periods, both in the embryo and during pupal metamorphosis, are developmental windows that are fundamentally activity-regulated, with an important role for shaping the network inhibition:excitation balance [1114]. For example, in Drosophila, transient correction of the network inhibition:excitation balance during the embryonic critical period effects a sustained rescue of seizure phenotypes otherwise caused by mutation of the single sodium channel [1114]. Experimentally, this model system allows reliable access to identified cells of known connectivity and function. Unlike mammals, Drosophila melanogaster, as an insect, is not warm-blooded, but poikilothermic, and is therefore vulnerable to external temperature variation. For the larval locomotor network, we found that increases in ambient temperature elevate network activity and, when applied during the embryonic critical period, phenocopy pharmacological and genetic activity manipulations. Thus, heat stress of 32°C, which is ecologically relevant to temperatures that Drosophila experience in the wild [15], represents an ecologically relevant stimulus with which to study critical period biology in this system.

Temperature affects biochemical reactions in general, and metabolic and mitochondrial processes in particular [16]. In Drosophila, heat stress causes a reversal of the flow of electrons within the mitochondrial electron transport chain (ETC), from Complex-II back to Complex-I. This so-called reverse electron transport (RET) leads to the production of reactive oxygen species (ROS) at Complex-I, due to a partial reduction of oxygen and the formation of superoxide anions. In Drosophila, as well as mouse, RET-generated ROS have been linked to improved stress resistance and life span [1721]. Here, we identified RET-based generation of ROS at mitochondrial Complex-I as a primary critical period signal that is necessary and sufficient for instructing changes in subsequent network development.

Downstream of mitochondrial ROS, we further identified the conserved Hypoxia-inducible factor 1 alpha (HIF-1α), of which Drosophila has only a single homologue, called similar (sima) [2224]. HIF-1α is a sensor for both oxygen and ROS levels, and is a regulator of metabolism [25]. During normal development, HIF-1α regulates many processes, including neurogenesis, angiogenesis, and cell survival [2628], and its dysregulation is linked to a range of cancers [29]. In summary, we report that critical period plasticity is principally enacted cell autonomously; with metabolic mitochondrial ROS as a primary signal that is transduced by the conserved HIF-1α pathway, instructing subsequent development of cellular excitable and synaptic properties, and thus network stability and behavioral output.

Results

Mitochondrial ROS are plasticity signals during the embryonic critical period

The primary signals and pathways that underlie critical period plasticity remain elusive. To identify these, we used the Drosophila larval neuromuscular junction (NMJ) as a proven experimental model [810]. As a means to induce a robust phenotype, we used transient exposure to 32°C heat stress during a defined embryonic critical period (occurring at 17–19 h after egg laying at 25°C). We have previously shown this manipulation to increase network excitation (i.e., increased action potential firing) phenocopying the effects of direct activity manipulations [30]. Specifically, embryos experiencing 32°C heat stress during their critical period show subsequent changes to larval development of the NMJ. This change is seemingly permanent and manifests as presynaptic terminal overgrowth (Fig 1A1C; S1 Data) and a decrease in the postsynaptic high-conductance glutamate receptor subunit, GluRIIA (Fig 1D; S1 Data), but not the lower conductance subunit, GluRIIB (Fig 1E; S1 Data).

Fig 1. Mitochondrial ROS generated by reverse electron transport in muscles is necessary for critical period heat stress to change NMJ development.

Fig 1

(A) Experimental paradigm. (B) 32°C during the critical period increases mitochondrial ROS production in muscles. Dot plots of mitochondrion-targeted ratiometric mito::roGFP2::Tsa2ΔCPΔCR ROS sensor in late stages at 25°C control temperature and 32°C. (C) Heat stress experienced during the embryonic critical period (32°C vs. 25°C control) leads to increased aCC NMJ terminal size and decreased postsynaptic GluRIIA, while not affecting subunit GluRIIB expression. Simultaneous genetic manipulation of muscle DA1 during embryonic stages only identifies mitochondrial ROS generated by reverse electron transport as necessary signals. “Control” indicates control genotype heterozygous for Oregon-R and DA1-GAL4. Larvae were reared at the control temperature of 25°C until the late wandering stage, 100 h after larval hatching (ALH). GluRIIA and GluRIIB subunits are displayed with lookup table “fire” to illustrate signal intensities (warmer colors indicating greater signal intensities). Scale bar = 20 µm. (D) Dot-plot quantification shows changes to aCC NMJ growth on its target muscle DA1, based on the standard measure of the number of boutons (swellings containing multiple presynaptic release sites/active zones). Data are shown with mean ± SEM, ANOVA, ***p < 0.0001, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (E) Dot-plot quantification shows changes in levels of the GluRIIA receptor subunit at aCC NMJs quantified in (C). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (F) Dot-plot quantification as in (D), but for the low conductance GluRIIB receptor subunit, which remains unaffected by these manipulations. See raw data in S1 Data.

The NMJ overgrowth phenotype mirrors the outcome following neuronal overactivation induced by ROS signaling, which we and others have previously studied [3032]. We therefore asked if mitochondrial ROS might be involved, and measured changes during the critical period following exposure to 32°C heat stress. Specifically, we expressed the ratiometric ROS sensor mito-roGFP2::Tsa2ΔCPΔCR, reported to have improved sensitivity and specificity over previously available tools [33] in all muscles using Mef2-GAL4. Indeed, quantification showed a significant increase in mitochondrial ROS in embryonic muscles following 32°C heat stress as compared to 25°C controls (Fig 1A and 1B; S1 Data).

To test if mitochondrial ROS were instructive critical period signals, we genetically dampened mitochondrial ROS levels in different ways, asking if this would suppress the critical period heat stress-induced NMJ phenotypes. We limited these genetic ROS manipulations to embryonic stages only, and, spatially, to a single muscle per half segment: the most dorsal muscle, dorsal acute 1 (DA1). Mis-expression of the mitochondria-targeted ROS scavengers, mito-Catalase or Perexiredoxin-3 (Prx3), in muscle DA1, cell-selectively rescued NMJ overgrowth (Fig 1D; S1 Data) and GluRIIA phenotypes that otherwise result from a critical period 32°C heat stress (Fig 1E; S1 Data). At the control temperature of 25°C, these same genetic manipulations show no such effects (Fig 1).

Mitochondria can generate ROS through various means, notably at Complex-I and Complex-III of the ETC. Heat stress has been shown to cause ROS generation at Complex-I in adult Drosophila, via reverse flow of electrons from Complex-II back to Complex-I, (termed ‘reverse electron transport’, or ‘RET’ [19]). To test if 32°C heat stress in embryos also causes RET-generated ROS in mitochondria, we cell-selectively knocked down a subunit of Complex-I, ND-75, required for RET-generated ROS. To complement this, in another manipulation we neutralized RET-generated ROS by mis-expression of alternative oxidase (AOX), an enzyme found in the inner mitochondrial membrane of many organisms, though not vertebrates or insects [3437]. In mitochondria, AOX acts as a bypass that transfers electrons from ubiquinone to oxygen, independent of Complexes-I and -III, thereby reducing the probability of electron backflow and subsequent ROS production by RET [19,3840]. When we targeted either of these genetic manipulations to muscle DA1 during embryogenesis, both rescued the critical period heat stress phenotype, as per NMJ bouton number and GluRII composition (Fig 1), as well as presynaptic active zone number (S2 Fig; S7 Data). Under control conditions of 25°C these manipulations were again neutral, leaving NMJ development unaffected (Figs 1 and S1; S6 Data). Together, our results show that during the embryonic critical period, RET-generated ROS at Complex-I are necessary for heat stress to effect a lasting change to NMJ development.

We next asked if ROS generated by Complex-I might also be sufficient to induce critical period plasticity in the absence of systemic heat stress. To test this, we cell-selectively induced RET-ROS generation at Complex-I by mis-expression of yeast NADH dehydrogenase 1 (Ndi1). Ndi1 transfers electrons from NADH to the CoQ pool. This can lead to the accumulation of reduced CoQ and transfer of electrons to Complex-I, causing RET and associated ROS generation [28]. Mis-expression of Ndi1in Drosophila has previously been shown to increase ROS production in a way indistinguishable from RET [20]. We found that transient embryonic mis-expression of Ndi1in muscle DA1, at the control temperature of 25°C, cell-selectively phenocopies the characteristic critical period heat stress phenotype of NMJ overgrowth (Figs 2A2C and S1; S2 and S6 Data), reduced GluRIIA levels (Figs 2E and S1) and an increase in active zone number (S2 Fig; S7 Data). Importantly, RET inhibition via AOX suppresses the effects of heat stress (Fig 1), while Ndi1-induced ROS at 25°C phenocopies the 32°C phenotype. To determine whether these manipulations act through a shared pathway, we combined Ndi1 mis-expression with critical period heat stress. This did not further enhance the phenotype, suggesting that temperature and Complex-I-derived ROS converge on the same signaling mechanism (Fig 2). Thus, our results show that RET-generated ROS, produced in mitochondria during the embryonic critical period, are both necessary and sufficient for inducing lasting change to NMJ development.

Fig 2. ROS by Complex-I, during the critical period, is sufficient to induce lasting changes to NMJ development.

Fig 2

(A) Experimental paradigm. (B) Induction of ROS at Complex-I via transient mis-expression of NDI1 in muscle DA1 during embryonic stages only phenocopies the effects that critical period heat stress has on subsequent NMJ development. “Control” indicates control genotype heterozygous for Oregon-R and DA1-GAL4. Larvae were reared at the control temperature of 25°C until the late wandering stage, 100 h ALH. GluRIIA and GluRIIB subunits are displayed with lookup table “fire” to illustrate signal intensities (warmer colors indicating greater signal intensities). Scale bar = 20 µm. (C) Dot-plot quantification shows changes to aCC NMJ growth on its target muscle DA1, based on the standard measure of the number of boutons (swellings containing multiple presynaptic release sites/active zones). Data are shown with mean ± SEM, ANOVA, ***p < 0.0001, ****p < 0.00001. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (D) Dot-plot quantification shows changes in levels of the GluRIIA receptor subunit at aCC NMJs quantified in C). Data are shown with mean ± SEM, ANOVA, **p < 0.001, ***p < 0.0001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (E) Dot-plot quantification as in (D), but for the low conductance GluRIIB receptor subunit, which remains unaffected by these manipulations. See raw data in S2 Data.

Mitochondrial RET-generated ROS are critical period signals for NMJ development

Synapse development requires close interactions between pre- and postsynaptic partners. Therefore, we tested the roles of mitochondrial ROS in a presynaptic motoneuron, termed ‘aCC’, whose NMJ forms on muscle DA1. We identified a transgenic line that transiently targets GAL4 expression to the aCC presynaptic motoneuron, here referred to as “aCC-GAL4[1]” (aka RN2-O-GAL4). Conveniently for critical period manipulations, GAL4 expression by this transgenic line is transient, limited to embryonic stages only, including the critical period. Using this aCC-GAL4[1] to express the mito-roGFP2::Tsa2ΔCPΔCR ROS sensor, we found that motoneurons, like muscles, increase ROS levels in mitochondria following 32°C heat stress during the critical period (Fig 3A and 3B). Next, we tested if ROS is also instructive signals in motoneurons during the critical period, as in muscles. We found that mis-expression of AOX in aCC motoneurons, to neutralize RET-induced ROS generation at Complex-I, was sufficient to rescue the NMJ overgrowth that normally results from embryonic 32°C heat stress (Fig 3A3D and 3F; S3 Data). Critical period-induced increases in NMJ size (as measured by either bouton number or area) led to a proportionate increase in the number of presynaptic active zones, the sites of vesicular neurotransmitter release, without affecting their overall density (Fig 3E and 3G; S3 Data). However, simultaneous embryonic mis-expression of AOX in the aCC motoneurons fully rescues active zone number (Fig 3D; S3 Data). Conversely, transient embryonic generation of RET-ROS in the aCC motoneurons, by mis-expression of Ndi1, was sufficient to cell-selectively mimic the critical period heat stress phenotype (Fig 3C3E; S3 Data). This phenotype was not exacerbated by simultaneous exposure to 32°C during the critical period (Fig 3). These findings demonstrate that NMJ phenotypes caused by critical period heat stress or ROS manipulations could be instigated as well as rescued through either presynaptic or postsynaptic processes. Interestingly, motoneuron manipulations during the critical period did not affect GluRII composition in muscles (S2 Fig; S7 Data).

Fig 3. ROS by RET in embryonic motoneurons instructs subsequent NMJ development.

Fig 3

(A) Experimental paradigm. (B) 32°C during the critical period increases mitochondrial ROS production in aCC motoneurons. Images generated by dividing the region of interest of 405 nm image by the same region obtained by 488 nm. Dot plots of mitochondrion-targeted ratiometric mito::roGFP2::Tsa2ΔCPΔCR ROS sensor in late stages at 25°C control temperature and 32°C. Scale bar = 10 µm. (C) Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress) and simultaneous, transient genetic manipulation of motoneuron aCC during embryonic stages only via aCC-GAL4[1] (akaRN2-O-GAL4). “Control” indicates control genotype heterozygous for Oregon-R and aCC-GAL4[1]. Larvae were reared at the control temperature of 25°C until the late wandering stage, 100 h ALH. Scale bar = 20 µm. (D) Dot-plot quantification shows changes to aCC NMJ growth on its target muscle DA1, based on the standard measure of the number of boutons (swellings containing multiple presynaptic release sites/active zones). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (E) Dot-plot quantification shows changes in the number of active zones at aCC NMJs quantified in (C). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (F) Dot-plot quantification shows changes in NMJ area. Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (G) Dot-plot quantification shows no changes in active zones density (active zone/μm2). Data are shown with mean ± SEM, ‘ns’ indicates statistical non-significance. Comparison with the control condition of 25°C throughout, genetically unmanipulated. See raw data in S3 Data.

Next, we tested the importance of developmental timing during which transient RET-generated ROS can cause such lasting change, i.e., whether mitochondrial ROS are indeed critical period-associated signals. To this end, we contrasted the effects of transient genetic manipulations of aCC motoneurons, during the late embryonic stage, including the critical period, versus identical manipulations after the critical period has closed (i.e., during the first larval stage). We used another, independently generated aCC expression line, “aCC-GAL4[2]” (aka GMR96G06-GAL4). aCC-GAL4[2] targets GAL4 to the aCC motoneuron from 2.5 h prior to critical period opening and then maintains expression during larval stages (S3 Fig) [41,42]. We achieved temporal control of GAL4 activity, independent of temperature, by combining aCC-GAL4[2] with the auxin-sensitive GAL4 inhibitor from the recently published “AGES” system [43]. To induce GAL4 activity during the embryonic critical period, we fed auxin to gravid females (thus introduced into eggs and embryos), while for later GAL4 expression, postcritical period closure, we fed auxin to freshly hatched larvae for 24 h. The results confirm that the timing of mitochondrial RET-generated ROS is important: only when cells generate mitochondrial ROS signals during the embryonic phase, including the critical period, do ROS cause lasting change, in keeping with this being a critical period of locomotor network development (S4 Fig; S8 Data).

The conserved HIF-1α pathway transduces the mitochondrial ROS signal to the nucleus

Then, we considered how mitochondrial ROS, triggered by heat stress, might effect change to developmental outcomes. HIF-1α is a transcription factor that plays a central role in cellular responses to hypoxia and other stresses, including oxidative stress [44]. Using endogenous HIF-1α tagged with GFP, we observed that both exposure to 32°C or mis-expression of Ndi1 lead to an increase in cytoplasmic HIF-1α::GFP and translocation to the nucleus, both in muscles and in motoneurons (Fig 4A). We asked if HIF-1α is indeed an instructive signal during the critical period. First, we knocked down the single Drosophila homologue of HIF-1α, sima, selectively in DA1 muscles while exposing embryos to 32°C during the critical period. This showed that HIF-1α is necessary to cause heat stress-induced critical period plasticity, e.g., NMJ overgrowth (Fig 4A4D; S4 Data) and a postsynaptic decrease of the GluRIIA receptor subunit (Fig 4E; S4 Data). Conversely, transient embryonic mis-expression of HIF-1α in DA1 muscles, at the control temperature of 25°C, is sufficient to induce these NMJ phenotypes (Fig 4). Under normoxic conditions, HIF-1α is continuously degraded by the proteasome, regulated via hydroxylation by the conserved prolyl hydroxylase, PHD (a single Drosophila homologue, Hph/fatiga). Hydroxylated HIF-1α is recognized by the E3 ligase complex substrate recognition subunit, von Hippel-Lindau (VHL), leading to HIF-1α ubiquitination and subsequent proteasomal degradation [45]. Mitochondrial ROS, including RET-generated, can inhibit PHD and thereby stabilize HIF-1α, also under normoxic conditions [23,4649]. To consolidate the role of HIF-1α signaling, we elicited transient stabilization of HIF-1α in embryonic muscle DA1 at the control temperature of 25°C, via RNAi knockdown of HIF-1α degradation machinery components: the prolyl hydroxylase, PHD, or the VHL ubiquitin ligase complex substrate recognition subunit. Both manipulations phenocopy the characteristic critical period-induced NMJ overgrowth phenotype (Fig 4).

Fig 4. HIF-1α is the signal downstream of ROS-RET.

Fig 4

(A) Experimental paradigm. (B) HIF-1α is stabilized and translocated to nucleus in muscles following 32°C or increase ROS by mitochondrial Complex-I (Ndi1 mis-expression) during the critical period. Images show GFP tagged endogenous HIF-1α and a nuclear expression of a red fluorescent protein. Scale bar = 100 µm. (C) Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress) and simultaneous genetic manipulation of muscle DA1 during embryonic stages only. “Control” indicates control genotype heterozygous for Oregon-R and DA1-GAL4. Larvae were reared at the control temperature of 25°C until the late wandering stage, 100 h ALH. GluRIIA subunit is displayed with a lookup table “fire” to illustrate signal intensities (warmer colors indicating greater signal intensities). Scale bar = 20 µm. (D) Dot-plot quantification shows changes to aCC NMJ growth on its target muscle DA1, based on the standard measure of the number of boutons (swellings containing multiple presynaptic release sites/active zones). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (E) Dot-plot quantification shows changes in levels of the GluRIIA receptor subunit at aCC NMJs quantified in (D). Data are shown with mean ± SEM, ANOVA, *p < 0.01, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. See raw data in S4 Data.

To further validate a working model of HIF-1α signaling as downstream of mitochondrial RET-generated ROS, we used a genetic epistasis experiment: inducing RET via mis-expression of Ndi1, while simultaneously knocking down the putative downstream acceptor, HIF-1α. We found that RET induction was no longer able to cause the expected changes in the larval NMJ when HIF-1α was knocked down at the same time (Fig 4). This supports a working model of information flow from mitochondrial RET-generated ROS to nuclear HIF-1α.

Testing this model of information flow in motoneurons, we selectively knocked down HIF-1α in motoneurons during critical period heat stress, which generates ROS via RET. This manipulation significantly suppressed both critical period heat stress-induced NMJ overgrowth and the concomitant increase in active zones. Thus, in motoneurons as in muscles, HIF-1α is required for critical period plasticity (Fig 5A and 5B; S5 Data). Given that all muscle-targeted manipulations of HIF-1α gain-of-function and stabilization by inhibition of degradation produced comparable phenotypes, we proceeded with the most effective approach: co-expression of HIF-1α gain-of-function together with RNAi against VHL. This was sufficient to fully recapitulate the heat-induced NMJ phenotype at the control temperature of 25°C (Fig 5A5D; S5 Data). We further challenged the working model by carrying out a genetic epistasis experiment, of inducing RET via mis-expression of Ndi1, while simultaneously knocking down HIF-1α. This did not have an impact at NMJ, reinforcing the conclusion that HIF-1α is the key effector downstream of mitochondrial ROS in mediating critical period plasticity.

Fig 5. ROS by RET and HIF-1α signaling during the embryonic critical period is necessary and sufficient to reduce motoneuron excitability and crawling behavior.

Fig 5

(A) Experimental paradigm. (B) HIF-1α is stabilized and translocated to nucleus in motoneurons following 32°C or increase ROS by mitochondrial Complex-I (Ndi1 mis-expression) during the critical period. Images show GFP tagged endogenous HIF-1α and a nuclear expression of a red fluorescent protein. Scale bar = 10 µm. (C) Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress) and simultaneous genetic manipulation of motoneuron aCC during embryonic stages only. “Control” indicates control genotype heterozygous for Oregon-R and aCC-GAL4[1]. Larvae were reared at the control temperature of 25°C until the late wandering stage, 100 h ALH. Scale bar = 20 µm. (D) Dot-plot quantification shows changes to aCC NMJ growth on its target muscle DA1, based on the standard measure of the number of boutons (swellings containing multiple presynaptic release sites/active zones). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (E) Dot-plot quantification shows changes in levels of the active zones in aCC motoneurons quantified in (D). Data are shown with mean ± SEM, ANOVA, *p < 0.01, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (F) aCC motoneuron excitability at the late third instar larval stage (i.e., action potential firing frequency triggered by current injected). Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress) and simultaneous transient genetic manipulation of aCC motoneuron during embryonic stages, via aCC-GAL4[1] (RN2-O-GAL4). “Control” indicates control genotype heterozygous for Oregon-R and aCC-GAL4[1]. Larvae were reared at the control temperature of 25°C until the late wandering stage, c. 100 h ALH. Control at 25°C vs. 32°C is significant at p = 0.0005; Control at 25°C vs. HIF-1α[GoF] is significant at p = 0.0015; Control at 25°C vs. AOX[GoF]2 at 32°C is not significant at p = 0.28. (G) Cell capacitance measurements as indicators of cell size show no significant differences between aCC motoneurons from specimens in (F). (H) Resting membrane potential, as an indicator for cell integrity, shows no significant differences between aCC motoneurons from specimens in (F). (I) Quantification of the first action potential amplitude (in mV) recorded from aCC motoneurons from specimens in (F). (J) Duration of first action potential fired (measured at 50% relative amplitude) shows no significant differences between aCC motoneurons from specimens in (F). (K) Crawling speed of third instar larvae (72 h ALH). Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress) and simultaneous genetic manipulation of CNS neurons during embryonic stages only. “Control” indicates control genotype heterozygous for Oregon-R and CNS-GAL4; Auxin-GAL80. Larvae were reared at the control temperature of 25°C. All animals, including “Control” are from gravid females fed with auxin. Each data point represents the crawling speed from an individual uninterrupted continuous forward crawl, n = specimen replicate number, up to three crawls assayed for each larva. Mean ± SEM, ANOVA, ns = not significant, ****p < 0.00001. See raw data in S5 Data.

Neuronal excitability and crawling speed are specified during the critical period by the mitochondrial ROS to HIF-1α signaling axis

Having thus far focused on the NMJ as a model synapse, we asked whether the structural changes that result from critical period plasticity have correlates in neuronal function and behavior [50]. To this end, we performed whole-cell patch clamp recordings from the aCC motoneuron at the late larval stage. These showed that embryonic heat stress of 32°C caused a lasting reduction in excitability (i.e., reduced capability to fire action potentials). Though cell size (assessed by membrane capacitance) and resting membrane potential were not affected. We also analyzed the first action potential fired by the lowest current injection: both the amplitude and the duration at half-maximum amplitude remained unchanged (Fig 5E5H; S5 Data). Thus, we are not yet able to propose a mechanism that underlies the observed reduction in excitability. This is further compounded by the fact that the action potential initiation zone is very far removed from the site of recordings [51].

At a network level, exposure to 32°C during the critical period results in reduced crawling speed at the late larval stage (Fig 5K; S5 Data). We find that mitochondrial RET-ROS and HIF-1α signaling are required and sufficient for this critical period-regulated change in locomotor network output. Transient induction of these signals, in all embryonic neurons, at the control temperature, phenocopies the reduction in crawling speed caused by critical period heat stress (Fig 5K; S5 Data). However, RET-ROS and HIF-1α signaling only have this lasting effect when induced during the embryonic phase, including the critical period window. In contrast, manipulations outside this plastic period, i.e., post critical period closing, are without effect (S5 Fig; S9 Data). Thus, the changes we observed following critical period heat stress, both centrally and at the NMJ, are sufficient to alter locomotor behavior.

To understand the underlying reason, we considered two scenarios that could explain reduced locomotor speed. First, embryonic exposure to 32°C may lead to larvae that are simply unable to crawl as fast as control animals. To test this, we acutely increased the temperature during the crawling assay and found that larvae responded with an increase in speed regardless of whether or not they had been exposed to heat stress as embryos. Moreover, under such conditions, critical period-manipulated animals could match the crawling speed of controls (S6A Fig; S10 Data). As a second hypothesis, embryonic heat stress during the critical period might ‘set’ a different locomotor network homeostatic setpoint. To test this, we maintained critical period-manipulated animals at 32°C for several hours. This results in animals reverting to their “default velocity”, including critical period-manipulated larvae, which reverted to the original reduced level, suggesting that the system is still capable of homeostasis: responding to change and returning to a pre-determined “default velocity” (S6B Fig; S10 Data). These observations suggest, that following critical period heat stress, the system is not simply damaged, but remains responsive to environmental temperature challenges, as in control animals. We speculate that following a critical period manipulation the locomotor network “default output” is positioned at a different set point or range.

We conclude that during the embryonic critical period, mitochondrial RET-generated ROS are instructive signals that are transmitted to the nucleus via the conserved transcriptional regulator, HIF-1α. Only when activated during the embryonic critical period, does this signaling pathway alter the subsequent development of the nervous system, leading to significant and permanent change in synaptic terminal growth, receptor composition, neuronal excitable properties, and network output, which manifests as changes in larval crawling behavior (Fig 6).

Fig 6. Working model.

Fig 6

(A) Heat stress during the critical period induces mitochondrial ROS, which leads to HIF-1α stabilization and its accumulation in the nucleus, where HIF-1α generates long-lasting changes that impact on (B) NMJ development: larger presynaptic terminals and less GluRIIA, as well as reductions in motoneuron excitability and, behaviorally, crawling speed. Figure A created using BioRender (https://www.biorender.com/).

Discussion

Working with a simplified experimental system, we have identified a metabolic signal, mitochondrial RET-generated ROS, as a primary critical period plasticity signal, that is transduced to the nucleus via the conserved HIF-1α signaling pathway. Moreover, we demonstrated that critical period-plasticity responses are fundamentally cell intrinsic, shown by virtue of targeting genetic manipulations, transiently, to single cells in vivo. This provides a first mechanistic explanation for how a perturbation experienced during a critical period effects a change in subsequent development, leading to altered neuronal properties and animal behavior.

Environmental cues are integrated during critical periods and direct nervous system development

It has long been known that early life experiences are formative, due to their potential to impart lasting influence on nervous system function [52,53]. Though critical periods have been most intensively studied in the context of activity-regulated sensory processing and cortical networks, it is possible that critical periods represent a fundamental process that is common to the development of nervous systems. For example, similar developmental windows of heightened vulnerability to disturbances have also been identified in mammalian locomotor systems [54], as well as in locomotor, visual, and olfactory systems of zebrafish [5559] and several insect species (ants, bees, and Drosophila [6064]). In fast-developing organisms, opening of a critical period correlates with the phase when network activity transitions from spontaneous and un-patterned to patterned activity, and in the case of the locomotor network of the Drosophila embryo lasts only 2 h [11,13,14,6567]. In mammalian systems, critical periods are much more drawn out, lasting weeks in mouse and up to several years in human [7]. Notwithstanding such differences, the neuronal activity-regulated processes of critical periods appear remarkably similar across species; from GABAergic signaling during critical periods of visual system and cortex development [59], to the importance of networks attaining an appropriate excitation:inhibition balance, also recently shown for the developing Drosophila locomotor network [12,13].

Here, we took advantage of the well-defined critical period of the Drosophila locomotor network [11,12]. As an ecologically relevant stimulus, we focused on temperature, specifically a 32°C heat stress, which we have previously shown phenocopies pharmacological or genetic activity manipulations, leading to similar lasting changes in subsequent neuronal development [50]. Using this experimental paradigm allowed us to identify a basic instructive signal that underlies critical period plasticity, namely mitochondrial ROS generated at Complex-I via reverse flow of electrons within the ETC. In mammals, mitochondrial RET-generated ROS has been associated with aging and expression of protective responses [19,68]. Whether this same signal is also responsible for the responses to developmental heat stress reported in ants [69], bees [62,70,71] or other parts and stages of the developing Drosophila nervous system [15,60], remains to be tested. Though we employed a stress signal in this study, to elicit a robust critical period plasticity response, temperature operates as a general environmental cue during nervous system development. Complementing this study, a cooler, non-stress temperature experienced during a pupal critical period of Drosophila nervous system development, has been reported to also lead to clear cellular effects. In that case, causing reduced filopodial dynamics and a consequent increase in synapse formation with concomitant changes in network connectivity and behavioral output [60,64].

Critical period effects at the level of single cells

One of the strengths of the experimental system we use is that it allows studying critical periods with single-cell resolution in vivo. Here, we demonstrated that critical period plasticity is primarily cell-autonomous. This simplifies establishing clear cause–effect relationships, which is more difficult with network-wide manipulations that are commonly used. Through cell-selective genetic manipulations of a single motoneuron or its target muscle, we phenotypically rescued individual cells from systemic heat stress manipulations. Thus, we demonstrated the cell-autonomous necessity of RET-generated mitochondrial ROS and HIF-1α signaling for heat stress-induced critical period plasticity. Conversely, we also showed cell-selective sufficiency of RET-ROS and HIF-1α signaling, which can phenocopy changes in NMJ development; though only when induced during the transient embryonic critical period, but not after.

We observed that muscle manipulations during late embryogenesis alter the composition of the postsynaptic receptors within the muscle. They also affect the development of the NMJ (increase in size and number of presynaptic release sites), which arises from an interplay between presynaptic motor axon terminal and postsynaptic muscle [72]. In contrast, motoneuron-targeted critical period manipulations do not affect the postsynaptic muscle GluRII composition. During embryogenesis, the body wall muscles become electrically active 2–3 h before the motoneuron is able to conduct action potentials and receive excitatory synaptic input [65,73,74]. This sequence of muscles maturing before presynaptic motoneurons would be compatible with muscles undergoing their critical period before motoneurons, and compatible with our observation of muscles having “decided” their GluRII composition before the motoneuron can influence other aspects of NMJ formation. From a perspective of the developing nervous system, it is well known that different regions undergo their respective critical periods in sequence [7]. This may be mirrored at the cellular level within a given region or network: different neuronal types might be integrated into circuits sequentially, for example, based on birth order or maturation speed, facilitating more reproducible and robust network assembly. In support of such a view, a recent study centered on developing mouse neocortex demonstrated the sequential maturation of two principal classes of inhibitory interneurons, and how this process directs network maturation [75].

Is the critical period a developmental window during which network setpoints are established?

Neuronal networks dynamically maintain stability and appropriate function around a given set point (or range), through homeostatic plasticity mechanisms [76,77]. How and when a set point is specified remains unknown. Previous evidence suggests that perturbations of activity during a critical period can alter the homeostatic set point, and thus lead to long-lasting changes in network function, including instability and epilepsy-like activity [11,78,79]. Thus, the focus has been on neuronal activity and its coordination as important processes in setting the set point [11,12,6567,80,81]. A challenge has been to relate the consequences of perturbing network activity, during a critical period, to specific changes in functional or behavioral outcomes. By working with a locomotor network, we found that heat stress at the cellular level, as a critical period perturbation, leads to reduced excitability of motoneurons. At the network level, this correlates with reduced crawling speed, which is seen as directly reflecting a change in output of the locomotor network. Importantly, changes in crawling speed only occurred following manipulations that spanned the embryonic critical period, and animals maintained the ability to increase their crawling speed. Moreover, network homeostasis appears to remain intact, as critical period-manipulated “slow” animals respond normally to acute upshifts in environmental temperature, but return to their slower “default” speed following chronic changes [30]. These observations are compatible with the possibility that the set point for locomotor network output is established during the critical period. Because this network is so well characterized and experimentally tractable, it now offers the opportunity to study the cellular mechanisms through which this is achieved.

Conserved mitochondrial-nuclear signaling directs critical period plasticity

Cells integrate internal (e.g., network activity) and external environmental cues (e.g., temperature) during a critical period [12]. The identification of an instructive signal that is intimately linked to mitochondrial metabolism significantly adds to, and complements, a large body of work that has focused on activity-regulated processes [1,2]. How a neuronal activity-regulated signal and mitochondrial metabolism interlink, especially during critical periods of development, will be important to understand. We had previously shown that mitochondrial and NADPH oxidase-generated ROS are required for activity-regulated neuronal plasticity outside the critical period [30,31,82]. Here, we identified a specific source of mitochondrial ROS, namely those generated at Complex-I via RET, which occurs when the coenzyme Q pool is reduced while the proton motif force is elevated [83]. A mitochondrial enzyme that is linked to the redox state of the coenzyme Q pool, dihydroorotate dehydrogenase (DHODH), has been identified as an important regulator of neuronal action potential firing rate, i.e., a determinant of the neuronal homeostatic setpoint [84,85]. Together with our findings here, we speculate that mitochondrial metabolism might be central to critical period plasticity, potentially encoding neuronal homeostatic setpoints.

How might a transient perturbation during a critical period lead to lasting change? HIF-1α, a transcriptional regulator involved in the cellular response to stress, is typically degraded under normal oxygen concentrations. The model we propose is that mitochondrial ROS, including those generated by RET, transiently inhibit the degradation of HIF-1α, thus allowing HIF-1α to accumulate and translocate into the nucleus [23,4649]. HIF-1α signaling is a known regulator of changes in cellular metabolism, for example, promoting aerobic glycolysis in a range of cancers [8688], but also during normal development, when rapid tissue growth is required [26,89,90]. HIF-1α achieves this through a combination of direct regulation of gene expression, notably of metabolic enzymes, and epigenetic modifications, partly executed through histone acetyl transferases that act as co-factors [9194]. In some forms of cancer, this is thought to create positive feedback loops, from elevated mitochondrial metabolism, to increased ROS production maintaining disinhibition of HIF-1α, which reinforces metabolic drive. It is therefore conceivable that transient HIF-1α signaling could initiate such metabolic-epigenetic feedback loops that maintain a change in metabolic state, and thus also of neuronal growth and functional properties.

Materials and Methods

Drosophila rearing and stocks

Drosophila stocks were kept on standard corn meal medium, at 25°C. For all experiments where crosses were necessary, stocks were selected and combined as a 1:2 ratio of GAL4-line males to UAS-line virgin females. The following Drosophila melanogaster strains were obtained from the Bloomington Stock Center: Oregon-R-P2 (BDSC_2376), Mef2-GAL4 (BDSC_27390), VNC-GAL4 (BDSC_40701), aCC-GAL4[2] (BDSC_39171), UAS-myr-GFP (32196), Hif-1α::GFP (aka sima-GFP) (BDSC_59821), UAS-RedStinger (BDSC_8546), UAS-Hif-1α[RNAi] (BDSC_26207), UAS-VHL[RNAi] (BDSC_50727), and UAS-PHD[RNAi] (BDSC_34717). The following strains were kindly provided as gifts: aCC-GAL4[1] (aka RN2-O-GAL4 or eve-GAL4.RN2-O) [95] and DA1-GAL4 (aka eve-GAL4.eme) [96] from Miki Fujioka; AtTIR1-T2A-AID-GAL80-AID inserted into attP-3B (from Tony Southall) [43], UAS-mito-catalase (from Alex Whitworth) [97], UAS-Prx3, UAS-ND75[RNAi], UAS-AOX[GoF]1 and UAS-AOX [GoF]2, UAS-Ndi1[GoF] (all from Alberto Sanz), and UAS-Hif-1α[GoF] (aka UAS-sima; from Joe Bateman). We generated the UAS-mito::roGFP2::Tsa2 ΔCPΔCR ROS sensor (see below). See Table 1 for more details of the strains used.

Table 1. Drosophila strains.

Drosophila strain genotype Additional information FlyBase identifiers Source
Oregon-R-P2 Wild Type FBst0002376 Bloomington Drosophila Stock Center; RRID: BDSC_2376
eve-GAL4[eme-GAL4] (aka SS2-C-GAL4; referred to as “DA1-GAL4”) GAL4 expression targeted to muscle DA1 during embryonic stages only FBtp0016392 Gift from Miki Fujioka; described in Han and Colleagues [96]
Mef2-GAL4 GAL4 expression targeted to all muscles FBst0027390 Bloomington Drosophila Stock Center; RRID: BDSC_27390
RN2-O-GAL4 (aka eve-GAL4.RN2}O; referred to as “aCC-GAL4 [1]”) GAL4 expression targeted to motorneurons aCC and RP2 during embryonic stages only FBti0038624 Gift from Miki Fujioka; generated by Fujioka and Colleagues [95]
GMR94G06-GAL4 (referred to as “aCC-GAL4 [2]”) GAL4 expression targeted to motorneuron aCC FBti0139793 Bloomington Drosophila Stock Center; RRID: BDSC_40701
GMR57C10-GAL4, AtTIR1-T2A-AID-Gal80-AID (aka n-synaptobrevin-GAL4; referred to as CNS-GAL4) GAL4 expression targeted to neurons FBti0137043 Bloomington Drosophila Stock Center; RRID: BDSC_39171
AtTIR1-T2A-AID-GAL80-AID (referred to as Auxin-GAL80) Ubiquitously expresses Gal80 sensitive to auxin FBti0216871 Bloomington Drosophila Stock Center; RRID: BDSC_92470; gift from Tony Southall
10xUAS-IVS-mito::roGFP2::Tsa2 ΔCPΔCR (referred to as UAS-mito::roGFP2::Tsa2 ΔCPΔCR) Mitochondrial targeted ROS sensor ------------ This paper: see methods
UAS-mito-catalase Mitochondrial-targeted catalase under UAS control FBtp0057181 Gift from Alex Whitworth; generated by Radyuk and Colleagues [97]
UAS-Prx3 Peroxiredoxin 3 under UAS control FBtp0057179 Gift from Alberto Sanz
UAS-ND75 [RNAi] Expresses dsRNA for RNAi of ND75 (FBgn0017566) under UAS control FBst0472606 Vienna Drosophila Resource Center #100733/KK
UAS-AOX [GoF]1 Expresses AOX from C. intestinalis under UAS control (allele F6) FBti0147237 Gift from Alberto Sanz
UAS-AOX [GoF]2 Expresses AOX from C. intestinalis under UAS control (allele F24) FBti0147239 Gift from Alberto Sanz
UAS-Ndi1 [GoF] Expresses Ndi1 from S.cerevisae under UAS control (allele A46 and B20) FBti0182574 and FBti0182575 Gift from Alberto Sanz
UAS-myrGFP Expresses GFP tagged with a myristoylation site under the control of UAS sequences FBst0032196 Bloomington Drosophila Stock Center; RRID: BDSC_32196
Hif-1-GFP GFP tagged at native locus of Hif-1 FBti0178346 Bloomington Drosophila Stock Center; RRID: BDSC_59821
UAS-nRFP Expresses a variant DsRed protein with a nuclear localization signal under the control of UAS FBti0040829 Bloomington Drosophila Stock Center; RRID: BDSC_8546
UAS-Hif-1 [RNAi] Expresses dsRNA for RNAi of sima (FBgn0266411) under UAS control FBti0115271 Bloomington Drosophila Stock Center; RRID: BDSC_26207
UAS-Hif-1 [GoF] (aka UAS-sima) Expresses Hif-1/sima (FBgn0266411) under UAS control FBti0076469 Bloomington Drosophila Stock Center; RRID: BDSC_9582
UAS-VHL [RNAi] Expresses dsRNA for RNAi of VHL (FBgn0041174) under UAS control FBti0158781 Bloomington Drosophila Stock Center; RRID: BDSC_50727
UAS-PHD [RNAi] Expresses dsRNA for RNAi of PHD (FBgn0264785) under UAS control FBti0140885 Bloomington Drosophila Stock Center; RRID: BDSC_34717

Critical period manipulations

Temperature manipulations.

Embryos were harvested from 0 to 6 h after egg lay at 25°C. Embryos were then shifted to their respective temperatures in incubators (25 or 32°C) for the remainder of embryogenesis. After all animals hatched, all were moved back to 25°C until assayed in late larval stages.

Auxin feeding.

Auxin (GK2088 1-Naphthaleneacetic acid potassium salt; Glentham) solution was mixed with melted molasses fly medium to achieve 5 mM auxin concentration and poured into petri dish plates. For embryonic exposure to auxin, adult flies in mixed-sex laying pots were fed on this food for 48 h prior to egg collection. For larval exposure, freshly hatched larvae were fed on this food for 24 h.

Dissections and immunocytochemistry

Flies were allowed to lay eggs on apple-juice agar-based medium overnight at 25°C. Larvae were then reared at 25°C on yeast paste, while avoiding over-crowding. Precise staging of the late wandering third instar stage was achieved by: a) checking that a proportion of animals from the same time-restricted egg lay had initiated pupariation; b) larvae had reached a certain size, and c) showed gut-clearance of food (yeast paste supplemented with Bromophenol Blue Sodium Salt (Sigma-Aldrich)). Larvae (of either sex) were dissected in Sorensen’s saline, fixed for 10 min at room temperature in Bouins fixative (Sigma-Aldrich), as previously detailed [31]. Wash solution was Sorensen’s saline containing 0.3% Triton X-100 (Sigma-Aldrich) and 0.25% BSA (Sigma-Aldrich). Primary antibodies, incubated overnight at 10°C, were: Goat-anti-HRP Alexa Fluor 488 (1:1000; Jackson ImmunoResearch Cat. No. 123-545-021), Rabbit-anti-GluRIIB (1:1000; gift from Mihaela Serpe [98]), Mouse anti-GluRIIA (1:100; Developmental Studies Hybridoma Bank, Cat. No. 8B4D2 (MH2B)); Mouse anti-BRP (1:100; Developmental Studies Hybridoma Bank, Cat. No. nc82); secondary antibodies, 2 h at room temperature: Donkey anti-Mouse StarRed (1:1000; Abberior, Cat. No. STRED-1001); and goat anti-Rabbit Atto594 (1:1000; Sigma-Aldrich Cat. No. 77671-1ML-F). Specimens were cleared in 80% glycerol, overnight at 4°C, then mounted in Mowiol.

Image acquisition and analysis

Specimens were imaged using an Olympus FV3000 point-scanning confocal, and a 60x/1.3N. A silicone oil immersion objective lens and FV31S-SW software. Confocal images were processed using ImageJ (to quantify GluRIIA and -B intensity) and Affinity Photo (Serif Ltd) to prepare figures. Bouton number of the aCC NMJ on muscle DA1 from segments A3-A5 was determined by counting every distinct spherical varicosity.

To quantify GluRIIA and -B fluorescence intensity, a threshold image for each channel was used to generate an outline selection, which included only the postsynaptic glutamate receptor clusters. This outline was then applied to the original image to define the regions of interest for analysis. Fluorescence intensities for GluRIIA and GluRIIB channels were measured within these outlines. Data were exported into Excel and the intensity values normalized to the negative control condition (25°C through all embryogenesis).

Generation of a ratiometric ROS reporter

The pJFRC12-10XUAS-IVS-mito-roGFP2::Tsa2ΔCPΔCR (for simplicity referred to as UAS-mito-roGFP2::Tsa2ΔCPΔCR) transgene was generated by Klenow assembly cloning (tinyurl.com/4r99uv8m). Briefly, we used pJFRC12-10XUAS-IVS-myr-GFP plasmid DNA, removed the coding sequence for myr::GFP using XhoI and XbaI, and replaced it with the mitochondrial targeting sequence from human COX8, derived from plasmid pMitoTimer [99], a gift from Zhen Yan (Addgene plasmid # 52659; http://n2t.net/addgene:52659; RRID: Addgene_52659); followed in frame by the coding sequence for roGFP2::Tsa2ΔCPΔCR from plasmid p415 TEF roGFP2‐Tsa2ΔCPΔCR [33], a gift from Tobias Dick (Addgene plasmid # 83239; http://n2t.net/addgene:83239; RRID: Addgene_83239). Transgenics were generated by microinjection via phiC31 integrase-mediated recombination [100] into landing site VK00037 [cytogenetic 22A3] by the Injection Service, Department of Genetics, University of Cambridge.

Fluorescence measurement of the ratiometric ROS reporter

Mef2-GAL4 was used for targeting expression of UAS-mito-roGFP2::Tsa2ΔCPΔCR in all muscles. Embryos from 25°C critical period (control) or 32°C critical period manipulations were dechorionated using bleach (3 min at room temperature). Late-stage embryos were selected, transferred onto double-sided sticky tape on glass slide, and covered with PBS, then imaged immediately with a Leica Stellaris point-scanning confocal, using a 63×/0.9NA (Leica) dipping objective lens. The ROS reporter was excited sequentially at 405 and 488 nm with emission detected at 500–600 nm. 16-bit images were acquired using Leica LAS X software and processed with ImageJ. Z-stack images were maximally projected. To remove fringing artifacts around bouton edges, 488 nm images were thresholded using the “Otsu” algorithm with values below threshold set to ‘not a number’, the mean value obtained from the same region of interest was taken for the 405 nm image and divided by the 488 nm image.

aCC-GAL4[1] was used to drive the expression of UAS-mito-roGFP2::Tsa2ΔCPΔCR in all aCC and RP2 motoneurons. Ventral nerve chords of freshly hatched (up to 2 h) larvae from 25°C critical period (control) or 32°C critical period manipulations were dissected in PBS-NEM (137 mM NaCl, 2.7 mM KCl, 10 mM Na2HPO4, 1.8 mM KH2PO4, 20 mM N-ethyl-maleimide (NEM), pH 7.4). Ventral nerve cord preparations were incubated for 5 min in PBS-NEM, then fixed for 8 min in 4% formaldehyde (in PBS-NEM). Specimens were washed three times in PBS-NEM and then equilibrated in 70% glycerol. Specimens were mounted in glycerol and imaged the same day on a Leica Stellaris point-scanning confocal, using a 40×/1.25NA (Leica) glycerol immersion objective lens. Analysis was carried out as per above.

aCC-GAL4[2] expression imaging

Embryos containing one copy of the aCC-GAL4[2] transgene and UAS-myr-GFP transgene reporter were dechorionated using bleach (3 min at room temperature). Specific development time points embryos were selected and fixed with 4% paraformaldehyde (Agar Scientific) in saline for 15 min, at room temperature. Following the standard procedures of washes in PBS containing 0.3% Triton X-100 and 0.25% (w/v) bovine serum albumin (Sigma-Aldrich), GAL4-directed expression of membrane-targeted GFP was visualized by FluoTag‐X4 single-domain anti-GFP antibody conjugated with Atto488 fluorophore (NanoTag Biotechnologies GmbH, Cat no: N0304-At488-L) and the longitudinal Fasciclin II-positive axon tracts via Mouse anti-Fasciclin II (1:10; Developmental Studies Hybridoma Bank, Cat. No. 1D4 anti-Fasciclin II) with the secondary Donkey anti-Mouse StarRed (1:1000; Abberior, cat. no. STRED-1001). Embryos were transferred onto double-sided sticky tape on glass slide for imaging.

For image GAL4 expression in larva, two time points were selected: 2 h after hatch and 72 h after hatch. Each larva was dissected in extracellular saline to isolate the CNS, using a hypodermic syringe needle (30 gauge; BD Microlance) as a scalpel. The CNS was then transferred onto poly-l-lysine-coated (Sigma-Aldrich) cover glass. Nerve cords were fixed and stained as per embryos, cleared in 70% glycerol, then mounted in Mowiol and sandwiched under a second cover glass, with thin aluminum foil strips used as spacers. Images were acquired using a Leica Stellaris point-scanning confocal with a Leica LAS X software and processed with ImageJ. Z-stack.

HIF-1α detection

We used a MiMIC-based RMCE line in which endogenous Hif-1α (aka Sima) was fused in-frame with GFP (see Table 1). To test the effect of Ndi1 [GoF], we crossed the Hif-1α-GFP stock with either DA1-GAL4 or aCC-GAL4[1] recombined with a nuclear-targeted red fluorescent protein reporter to mark manipulated cells. To detect expression in muscles, embryos from 25°C critical period (control) or 32°C critical period or Ndi1 [GoF] manipulations were dechorionated and mounted in saline as per above shortly after critical period closure (tracheal filling), then imaged with a CSU-22 spinning disc confocal, using an Olympus 63×/0.8NA dipping objective lens. For motoneuron analysis of Hif-1α expression, ventral nerve chords of freshly hatched larvae from 25°C critical period (control), 32°C critical period or Ndi1 [GoF] manipulations were dissected in Sorensen’s saline and transferred to a cover glass coated with poly-l-lysine (Sigma-Aldrich). Ventral nerve cord preparations were then fixed for 8 min in 4% formaldehyde (in PBS). Specimens were washed three times in PBS and then equilibrated in 70% glycerol and imaged the same day on a Leica Stellaris point-scanning confocal, using a 40×/NA1.25 (Leica) glycerol immersion objective lens.

Electrophysiology

Third instar (L3) larvae (of either sex) were dissected in a dish containing standard saline (135 mM NaCl (Fisher Scientific), 5 mM KCl (Fisher Scientific), 4 mM MgCl2·6H2O (Sigma-Aldrich), 2 mM CaCl2·2H2O (Fisher Scientific), 5 mM TES (Sigma-Aldrich), 36 mM sucrose (Fisher Scientific), pH 7.15) to remove the ventral nerve cord and brain lobes (CNS). The isolated CNS was then transferred to a droplet of external saline containing 200 µM mecamylamine (Sigma-Aldrich) to block postsynaptic nACh receptors in order to synaptically isolate motor neurons. CNSs were laid flat (dorsal side up) and glued (GLUture Topical Tissue Adhesive; World Precision Instruments USA) to a Sylgard-coated cover slip (1–2 mm depth of cured SYLGARD Elastomer (Dow-Corning USA) on a 22 × 22 mm square coverslip. Preparations were placed on a glass slide under a microscope (Olympus BX51-WI), viewed using a 60× water-dipping lens. To access nerve cell somata, 1% protease (Streptomyces griseus, Type XIV, Sigma-Aldrich, in external saline) contained within a wide-bore glass pipette (GC100TF-10; Harvard Apparatus UK, ~10 µm opening) was applied to abdominal segments, roughly between A5-A2 [101] to remove overlaying glia. Motoneurons were identified by anatomical position and relative cell size, with aCC being positioned close to the midline and containing both an ipsilateral and contralateral projection. Whole cell patch clamp recordings were made using borosilicate glass pipettes (GC100F-10, Harvard Apparatus) that were fire polished to resistances of 10–15 MΩ when filled with intracellular saline (140 mM potassium-d-gluconate (Sigma-Aldrich), 2m M MgCl2·6H2O (Sigma-Aldrich), 2 mM EGTA (Sigma-Aldrich), 5 mM KCl (Fisher Scientific), and 20 mM HEPES (Sigma-Aldrich), (pH 7.4). Input resistance was measured in the ‘whole cell’ configuration, and only cells that had an input resistance ≥0.5 GΩ were used for experiments. Cell capacitance and break-in resting membrane potential were measured for each cell recorded; only cells with a membrane potential upon break in of <−40 mV were analyzed. Data for current step recordings were captured using a Multiclamp 700B amplifier (Molecular Devices) controlled by pCLAMP (version 10.7.0.3), via an analogue-to-digital converter (Digidata 1440A, Molecular Devices). Recordings were sampled at 20 kHz and filtered online at 10 kHz. Once patched, neurons were brought to a membrane potential of −60 mV using current injection. Each recording consisted of 20 × 4pA (500 ms) current steps, including an initial negative step, giving a range of −4 to +72pA. The number of action potentials was counted at each step and plotted against injected current. Cell capacitance and membrane potential were measured between conditions to ensure that any observed differences in excitability were not due to differences in either cell size or resting state.

Larval crawling analysis

At the mid-third instar stage, 72 h after larval hatching (ALH), larvae were rinsed in water and placed inside a 24 cm × 24 cm crawling arena with a base of 5 mm thick 0.8% agarose gel (Bacto Agar), situated inside an incubator. Temperature was maintained at 25 ± 0.5°C, reported via a temperature probe in the agar medium. Humidity was kept constant. Larval crawling was recorded using a frustrated total internal reflection-based imaging method (FIM) in conjunction with the tracking software FIMTrack [102,103] using a Basler acA2040-180km CMOS camera with a 16 mm KOWA-IJM3sHC.SW-VIS-NIR Lens controlled by Pylon (by Basler) and Streampix (v.6) software (by NorPix). Larvae were recorded for 20 min at five frames per second.

Recordings were split into four 5-min sections with the first section not tracked and analyzed as larvae are acclimatizing to the crawling arena. The remaining three sections were used to analyze crawling speed, with each larva sampled at most once per section. Crawling speed was calculated using FIMTrack software choosing crawling tracks that showed periods of uninterrupted forward crawling devoid of pauses, turning, or collision events. The distance traveled was determined via the FIMTrack output.

Statistical analysis

Statistical analyses were carried out using GraphPad Prism Software (Version 10.1.2). Datasets were tested for normal distribution with the Shapiro-Wilk Test. Normally distributed data were then tested with student t test (for pairwise comparison). Normally distributed analysis for more than two groups was done with a one-way ANOVA and post hoc tested with a Tukey multiple comparison test. Non-normally distributed data sets of two groups were tested with Mann–Whitney U Test (pairwise comparison) and datasets with more than two groups were tested with a Kruskal–Wallis ANOVA and post hoc tested with a Dunn’s multiple post hoc comparison test. For whole cell electrophysiology in aCC neurons, linear regression analysis was used to compare the intercepts between conditions, using a Bonferroni correction for multiple comparisons, resulting in an adjusted significance threshold of p < 0.0083. This analysis was restricted to the linear portions of the input-output curves. For all datasets mean and standard error of mean (SEM) are shown. Significance levels were *p < 0.05; **p < 0.01; ***p < 0.001; ****p < 0.0001.

Supporting information

S1 Fig. Control of UAS-lines crossed to wild type Oregon-R.

(A) Dot-plot quantification shows changes to aCC NMJ growth on its target muscle DA1, based on the standard measure of the number of boutons (swellings containing multiple presynaptic release sites/active zones). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the wild type of condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with wild type exposed to 32°C heat stress during the embryonic critical period. “Wild type” is Oregon-R. (B) Dot-plot quantification shows changes in levels of the GluRIIA receptor subunit at aCC NMJs quantified in C). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the wild type condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with wild type exposed to 32°C heat stress during the embryonic critical period. “Wild type” is Oregon-R. See raw data in S6 Data.

(TIFF)

pbio.3003338.s001.tiff (419.5KB, tiff)
S2 Fig. Muscles influence motoneuron at NMJ during the critical period, but motoneuron do not affect GluR composition in muscles.

(A) Heat stress experienced during the embryonic critical period (32°C vs. 25°C control) leads to decreased postsynaptic GluRIIA. Simultaneous genetic manipulation of aCC motoneuron during embryonic stages does not affect GluRIIA in muscles. “Control” indicates control genotype heterozygous for Oregon-R and aCC-GAL4[1]. Larvae were reared at the control temperature of 25°C until the late wandering stage, 100 h after larval hatching (ALH). Data are shown with mean ± SEM, ANOVA, **p < 0.001, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (B) Heat stress experienced during the embryonic critical period (32°C vs. 25°C control) leads to increased presynaptic active zone number. Simultaneous genetic manipulation of muscle DA1 during embryonic stages suggests that ROS by Complex-I is necessary and sufficient to induce active zones changes presynaptically. “Control” indicates control genotype heterozygous for Oregon-R and DA1-GAL4. Larvae were reared at the control temperature of 25°C until the late wandering stage, 100 h after larval hatching (ALH). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. See raw data in S7 Data.

(TIFF)

pbio.3003338.s002.tiff (478.6KB, tiff)
S3 Fig. Characterization of aCC-GAL4 [2] expression.

Expression starts around 13 h after egg lay (AEL) in few sporadic cells, not including the aCC motoneuron and mostly limited to the head region. First aCC motoneurons show GAL4 expression as of 14.5 h AEL, with more aCC motoneurons expressing from 15 h AEL onwards (i.e., 2 h prior to critical period opening). Concomitantly, GAL4 expression in other cells disappears. From laraval hatching onwards, segmentally repeated expression in all aCC motoneurons is maintained until at least 72 h after larval hatching (mid-3rd instar stage). Scale bar = 100 µm.

(TIFF)

pbio.3003338.s003.tiff (898.2KB, tiff)
S4 Fig. RET-ROS and HIF-1α signaling has lasting impact on NMJ only during the critical period.

(A) Experimental paradigm. (B) Genetic manipulation of motoneuron aCC during the critical period versus after critical period closure. “Control” indicates control genotype heterozygous for Oregon-R and aCC-GAL4[2]; Auxin-GAL80. Larvae were reared at the control temperature of 25oC until the late wandering stage, 100 h ALH. Scale bar = 20 µm. (C) Dot-plot quantification shows changes to aCC NMJ growth on its target muscle DA1, based on the standard measure of the number of boutons (swellings containing multiple presynaptic release sites/active zones). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. (D) Dot-plot quantification shows changes in the number of active zones at aCC NMJs quantified in C). Data are shown with mean ± SEM, ANOVA, ***p < 0.0001, ****p < 0.00001, ‘ns’ indicates statistical non-significance. See raw data in S8 Data.

(TIFF)

pbio.3003338.s004.tiff (775.3KB, tiff)
S5 Fig. RET-ROS and HIF-1α signaling have no lasting impact on behavior after the critical period.

(A) Experimental paradigm. (B) Crawling speed of third instar larvae (72 h after larval hatching (ALH)). Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress) and simultaneous genetic manipulation of all neurons during first instar stage. “Control” indicates control genotype heterozygous for Oregon-R and CNS-GAL4; Auxin-GAL80. Larvae were reared at the control temperature of 25°C until 72 h ALH. Each data point represents crawling speed from an individual uninterrupted continuous forward crawl, n = specimen replicate number, up to three crawls assayed for each larva. Mean ± SEM, ANOVA, ns = not significant, ****p < 0.00001. See raw in S9 Data.

(TIFF)

pbio.3003338.s005.tiff (303.3KB, tiff)
S6 Fig. Homeostatic responses to environmental change are maintained after 32°C critical period.

(A) Acute shift of temperature. Crawling speed of third instar larvae (72 h after larval hatching (ALH)). Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress); temperature experienced during the larva rearing (25oC control vs. 32°C heat stress) and control (25°C) versus acute shift of temperature during the crawling assay (32oC). Genotype is Oregon-R. Each data point represents crawling speed from an individual uninterrupted continuous forward crawl, n = specimen replicate number, up to three crawls assayed for each larva. Mean ± SEM, ANOVA, ns = not significant, ****p < 0.00001. (B) Chronic adaptation to the heat stress. Crawling speed of third instar larvae (72 h after larval hatching (ALH)). Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress); temperature experienced during the larva rearing and during the crawling assay (25°C control vs. 32°C chronic adaptation). Oregon-R. Each data point represents crawling speed from an individual uninterrupted continuous forward crawl, n = specimen replicate number, up to three crawls assayed for each larva. Mean ± SEM, ANOVA, ns = not significant. See raw data in S10 Data.

(TIFF)

pbio.3003338.s006.tiff (336.4KB, tiff)
S1 Data. Raw date of results in Fig 1.

(XLSX)

pbio.3003338.s007.xlsx (23.1KB, xlsx)
S2 Data. Raw date of results in Fig 2.

(XLSX)

pbio.3003338.s008.xlsx (14.1KB, xlsx)
S3 Data. Raw date of results in Fig 3.

(XLSX)

pbio.3003338.s009.xlsx (19.3KB, xlsx)
S4 Data. Raw date of results in Fig 4.

(XLSX)

pbio.3003338.s010.xlsx (16.5KB, xlsx)
S5 Data. Raw date of results in Fig 5.

(XLSX)

pbio.3003338.s011.xlsx (22.1KB, xlsx)
S6 Data. Raw date of results in S1 Fig.

(XLSX)

pbio.3003338.s012.xlsx (14.8KB, xlsx)
S7 Data. Raw date of results in S2 Fig.

(XLSX)

pbio.3003338.s013.xlsx (12KB, xlsx)
S8 Data. Raw date of results in S4 Fig.

(XLSX)

pbio.3003338.s014.xlsx (13.2KB, xlsx)
S9 Data. Raw date of results in S5 Fig.

(XLSX)

pbio.3003338.s015.xlsx (13.2KB, xlsx)
S10 Data. Raw date of results in S6 Fig.

(XLSX)

pbio.3003338.s016.xlsx (16.1KB, xlsx)

Acknowledgments

The authors are grateful to Mihaela Serpe for the Rabbit polyclonal anti-GluRIIB antiserum.

The authors are grateful to Tony Southall, Alex Whitworth, Joe Bateman, and Alberto Sanz for generously providing fly stocks. Stocks obtained from the Bloomington Drosophila Stock Center (NIH P40OD018537) were used in this study. The nc82 and 1D4 monoclonal antibodies were obtained from the Developmental Studies Hybridoma Bank, created by the NICHD of the NIH and maintained at The University of Iowa, Department of Biology, Iowa City, IA 52242, USA. The work benefited from the Imaging Facility, Department of Zoology, supported by Matt Wayland, and funds from a Wellcome Trust Equipment Grant (WT079204) with contributions by the Sir Isaac Newton Trust in Cambridge, including Research Grant [18.07ii(c)]. Work on this project benefited from the Manchester Fly Facility, established through funds from the University of Manchester and the Wellcome Trust (Grant 087742/Z/08/Z).

Abbreviations

ALH

after larval hatching

AOX

alternative oxidase

DA1

dorsal acute 1

DHODH

dihydroorotate dehydrogenase

ETC

electron transport chain

HIF-1α

hypoxia-inducible factor

Ndi1

NADH dehydrogenase 1

NMJ

neuromuscular junction

Prx3

Perexiredoxin-3

RET

reverse electron transport

ROS

reactive oxygen species

VHL

von Hippel-Lindau

Data Availability

All relevant data are within the paper and its Supporting information files.

Funding Statement

D.S.-C. was supported by the European Molecular Biology Organization (EMBO) with a long-term EMBO fellowship (ALTF 62-2021). This work was supported by a Joint Wellcome Trust Investigator Award to R.A.B. and M.L. (217099/Z/19/Z) and funding from the Biotechnology and Biological Sciences Research Council (BBSRC) to M.L. (BB/V014943/1). The sponsors or funders had no role in the study design, data collection and analysis, decision to publish, or preparation of the manuscript.

References

  • 1.Hensch TK. Critical period regulation. Annu Rev Neurosci. 2004;27:549–79. doi: 10.1146/annurev.neuro.27.070203.144327 [DOI] [PubMed] [Google Scholar]
  • 2.Hensch TK. Critical period plasticity in local cortical circuits. Nat Rev Neurosci. 2005;6(11):877–88. doi: 10.1038/nrn1787 [DOI] [PubMed] [Google Scholar]
  • 3.Hensch TK, Quinlan EM. Critical periods in amblyopia. Vis Neurosci. 2018;35:E014. doi: 10.1017/S0952523817000219 [DOI] [PubMed] [Google Scholar]
  • 4.Do KQ, Cuenod M, Hensch TK. Targeting oxidative stress and aberrant critical period plasticity in the developmental trajectory to schizophrenia. Schizophr Bull. 2015;41(4):835–46. doi: 10.1093/schbul/sbv065 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Steullet P, Cabungcal J-H, Coyle J, Didriksen M, Gill K, Grace AA, et al. Oxidative stress-driven parvalbumin interneuron impairment as a common mechanism in models of schizophrenia. Mol Psychiatry. 2017;22(7):936–43. doi: 10.1038/mp.2017.47 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Sakai A, Sugiyama S. Experience-dependent transcriptional regulation in juvenile brain development. Dev Growth Differ. 2018;60(8):473–82. doi: 10.1111/dgd.12571 [DOI] [PubMed] [Google Scholar]
  • 7.Reh RK, Dias BG, Nelson CA 3rd, Kaufer D, Werker JF, Kolb B, et al. Critical period regulation across multiple timescales. Proc Natl Acad Sci U S A. 2020;117(38):23242–51. doi: 10.1073/pnas.1820836117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Collins CA, Wairkar YP, Johnson SL, DiAntonio A. Highwire restrains synaptic growth by attenuating a MAP kinase signal. Neuron. 2006;51(1):57–69. doi: 10.1016/j.neuron.2006.05.026 [DOI] [PubMed] [Google Scholar]
  • 9.Shen W, Ganetzky B. Autophagy promotes synapse development in Drosophila. J Cell Biol. 2009;187(1):71–9. doi: 10.1083/jcb.200907109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Frank CA, Wang X, Collins CA, Rodal AA, Yuan Q, Verstreken P, et al. New approaches for studying synaptic development, function, and plasticity using Drosophila as a model system. J Neurosci. 2013;33(45):17560–8. doi: 10.1523/JNEUROSCI.3261-13.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Giachello CNG, Baines RA. Inappropriate neural activity during a sensitive period in embryogenesis results in persistent seizure-like behavior. Curr Biol. 2015;25(22):2964–8. doi: 10.1016/j.cub.2015.09.040 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Hunter I, Coulson B, Pettini T, Davies JJ, Parkin J, Landgraf M, et al. Balance of activity during a critical period tunes a developing network. Elife. 2024;12:RP91599. doi: 10.7554/eLife.91599 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Carreira-Rosario A, York RA, Choi M, Doe CQ, Clandinin TR. Mechanosensory input during circuit formation shapes Drosophila motor behavior through patterned spontaneous network activity. Curr Biol. 2021;31(23):5341-5349.e4. doi: 10.1016/j.cub.2021.08.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Zeng X, Komanome Y, Kawasaki T, Inada K, Jonaitis J, Pulver SR, et al. An electrically coupled pioneer circuit enables motor development via proprioceptive feedback in Drosophila embryos. Curr Biol. 2021;31(23):5327-5340.e5. doi: 10.1016/j.cub.2021.10.005 [DOI] [PubMed] [Google Scholar]
  • 15.Williamson M, Mitchell A, Condron B. Birth temperature followed by a visual critical period determines cooperative group membership. J Comp Physiol A Neuroethol Sens Neural Behav Physiol. 2021;207(6):739–46. doi: 10.1007/s00359-021-01512-3 [DOI] [PubMed] [Google Scholar]
  • 16.Somero GN, Lockwood BL, Tomanek L. Biochemical adaptation: response to environmental challenges from life’s origins to the Anthropocene. Sinauer Associates, Inc. Publishers. 2017. [Google Scholar]
  • 17.Scialò F, Sriram A, Fernández-Ayala D, Gubina N, Lõhmus M, Nelson G, et al. Mitochondrial ROS produced via reverse electron transport extend animal lifespan. Cell Metab. 2016;23(4):725–34. doi: 10.1016/j.cmet.2016.03.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Scialò F, Fernández-Ayala DJ, Sanz A. Role of mitochondrial reverse electron transport in ROS signaling: potential roles in health and disease. Front Physiol. 2017;8:428. doi: 10.3389/fphys.2017.00428 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Scialò F, Sriram A, Stefanatos R, Spriggs RV, Loh SHY, Martins LM, et al. Mitochondrial complex I derived ROS regulate stress adaptation in Drosophila melanogaster. Redox Biol. 2020;32:101450. doi: 10.1016/j.redox.2020.101450 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Sanz A, Soikkeli M, Portero-Otín M, Wilson A, Kemppainen E, McIlroy G, et al. Expression of the yeast NADH dehydrogenase Ndi1 in Drosophila confers increased lifespan independently of dietary restriction. Proc Natl Acad Sci U S A. 2010;107(20):9105–10. doi: 10.1073/pnas.0911539107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Sanz A, Stefanatos R, McIlroy G. Production of reactive oxygen species by the mitochondrial electron transport chain in Drosophila melanogaster. J Bioenerg Biomembr. 2010;42(2):135–42. doi: 10.1007/s10863-010-9281-z [DOI] [PubMed] [Google Scholar]
  • 22.Chandel NS, McClintock DS, Feliciano CE, Wood TM, Melendez JA, Rodriguez AM, et al. Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1alpha during hypoxia: a mechanism of O2 sensing. J Biol Chem. 2000;275(33):25130–8. doi: 10.1074/jbc.M001914200 [DOI] [PubMed] [Google Scholar]
  • 23.Movafagh S, Crook S, Vo K. Regulation of hypoxia-inducible factor-1a by reactive oxygen species: new developments in an old debate. J Cell Biochem. 2015;116(5):696–703. doi: 10.1002/jcb.25074 [DOI] [PubMed] [Google Scholar]
  • 24.Qutub AA, Popel AS. Reactive oxygen species regulate hypoxia-inducible factor 1alpha differentially in cancer and ischemia. Mol Cell Biol. 2008;28(16):5106–19. doi: 10.1128/MCB.00060-08 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Huang X, Zhao L, Peng R. Hypoxia-inducible factor 1 and mitochondria: an intimate connection. Biomolecules. 2022;13(1):50. doi: 10.3390/biom13010050 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Tomita S, Ueno M, Sakamoto M, Kitahama Y, Ueki M, Maekawa N, et al. Defective brain development in mice lacking the Hif-1alpha gene in neural cells. Mol Cell Biol. 2003;23(19):6739–49. doi: 10.1128/MCB.23.19.6739-6749.2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Semenza GL. Regulation of physiological responses to continuous and intermittent hypoxia by hypoxia-inducible factor 1. Exp Physiol. 2006;91(5):803–6. doi: 10.1113/expphysiol.2006.033498 [DOI] [PubMed] [Google Scholar]
  • 28.Milosevic J, Maisel M, Wegner F, Leuchtenberger J, Wenger RH, Gerlach M, et al. Lack of hypoxia-inducible factor-1 alpha impairs midbrain neural precursor cells involving vascular endothelial growth factor signaling. J Neurosci. 2007;27(2):412–21. doi: 10.1523/JNEUROSCI.2482-06.2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Iacobini C, Vitale M, Pugliese G, Menini S. The “sweet” path to cancer: focus on cellular glucose metabolism. Front Oncol. 2023;13:1202093. doi: 10.3389/fonc.2023.1202093 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Oswald MC, Brooks PS, Zwart MF, Mukherjee A, West RJ, Giachello CN, et al. Reactive oxygen species regulate activity-dependent neuronal plasticity in Drosophila. Elife. 2018;7:e39393. doi: 10.7554/eLife.39393 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Sobrido-Cameán D, Oswald MCW, Bailey DMD, Mukherjee A, Landgraf M. Activity-regulated growth of motoneurons at the neuromuscular junction is mediated by NADPH oxidases. Front Cell Neurosci. 2023;16:1106593. doi: 10.3389/fncel.2022.1106593 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Stavrovskaya I, Morin BK, Madamba S, Alexander C, Romano A, Alam S, et al. Mitochondrial ROS modulate presynaptic plasticity in the Drosophila neuromuscular junction. Redox Biol. 2025;79:103474. doi: 10.1016/j.redox.2024.103474 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Morgan B, Van Laer K, Owusu TNE, Ezeriņa D, Pastor-Flores D, Amponsah PS, et al. Real-time monitoring of basal H2O2 levels with peroxiredoxin-based probes. Nat Chem Biol. 2016;12(6):437–43. doi: 10.1038/nchembio.2067 [DOI] [PubMed] [Google Scholar]
  • 34.McDonald A, Vanlerberghe G. Branched mitochondrial electron transport in the Animalia: presence of alternative oxidase in several animal phyla. IUBMB Life. 2004;56(6):333–41. doi: 10.1080/1521-6540400000876 [DOI] [PubMed] [Google Scholar]
  • 35.McDonald AE, Vanlerberghe GC, Staples JF. Alternative oxidase in animals: unique characteristics and taxonomic distribution. J Exp Biol. 2009;212(Pt 16):2627–34. doi: 10.1242/jeb.032151 [DOI] [PubMed] [Google Scholar]
  • 36.Matus-Ortega MG, Salmerón-Santiago KG, Flores-Herrera O, Guerra-Sánchez G, Martínez F, Rendón JL, et al. The alternative NADH dehydrogenase is present in mitochondria of some animal taxa. Comp Biochem Physiol Part D Genomics Proteomics. 2011;6(3):256–63. doi: 10.1016/j.cbd.2011.05.002 [DOI] [PubMed] [Google Scholar]
  • 37.Weaver RJ, McDonald AE. Mitochondrial alternative oxidase across the tree of life: presence, absence, and putative cases of lateral gene transfer. Biochim Biophys Acta Bioenerg. 2023;1864(4):149003. doi: 10.1016/j.bbabio.2023.149003 [DOI] [PubMed] [Google Scholar]
  • 38.Maxwell DP, Wang Y, McIntosh L. The alternative oxidase lowers mitochondrial reactive oxygen production in plant cells. Proc Natl Acad Sci U S A. 1999;96(14):8271–6. doi: 10.1073/pnas.96.14.8271 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Fernandez-Ayala DJM, Sanz A, Vartiainen S, Kemppainen KK, Babusiak M, Mustalahti E, et al. Expression of the Ciona intestinalis alternative oxidase (AOX) in Drosophila complements defects in mitochondrial oxidative phosphorylation. Cell Metab. 2009;9(5):449–60. doi: 10.1016/j.cmet.2009.03.004 [DOI] [PubMed] [Google Scholar]
  • 40.El-Khoury R, Dufour E, Rak M, Ramanantsoa N, Grandchamp N, Csaba Z, et al. Alternative oxidase expression in the mouse enables bypassing cytochrome c oxidase blockade and limits mitochondrial ROS overproduction. PLoS Genet. 2013;9(1):e1003182. doi: 10.1371/journal.pgen.1003182 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Pfeiffer BD, Jenett A, Hammonds AS, Ngo T-TB, Misra S, Murphy C, et al. Tools for neuroanatomy and neurogenetics in Drosophila. Proc Natl Acad Sci U S A. 2008;105(28):9715–20. doi: 10.1073/pnas.0803697105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Liu K, Jones S, Minis A, Rodriguez J, Molina H, Steller H. PI31 is an adaptor protein for proteasome transport in axons and required for synaptic development. Dev Cell. 2019;50(4):509-524.e10. doi: 10.1016/j.devcel.2019.06.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.McClure CD, Hassan A, Aughey GN, Butt K, Estacio-Gómez A, Duggal A, et al. An auxin-inducible, GAL4-compatible, gene expression system for Drosophila. Elife. 2022;11:e67598. doi: 10.7554/eLife.67598 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Weidemann A, Johnson RS. Biology of HIF-1alpha. Cell Death Differ. 2008;15(4):621–7. doi: 10.1038/cdd.2008.12 [DOI] [PubMed] [Google Scholar]
  • 45.Strowitzki MJ, Cummins EP, Taylor CT. Protein hydroxylation by hypoxia-inducible factor (HIF) hydroxylases: unique or ubiquitous? Cells. 2019;8(5):384. doi: 10.3390/cells8050384 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Chandel NS, Maltepe E, Goldwasser E, Mathieu CE, Simon MC, Schumacker PT. Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc Natl Acad Sci U S A. 1998;95(20):11715–20. doi: 10.1073/pnas.95.20.11715 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Brunelle JK, Bell EL, Quesada NM, Vercauteren K, Tiranti V, Zeviani M, et al. Oxygen sensing requires mitochondrial ROS but not oxidative phosphorylation. Cell Metab. 2005;1(6):409–14. doi: 10.1016/j.cmet.2005.05.002 [DOI] [PubMed] [Google Scholar]
  • 48.Guzy RD, Hoyos B, Robin E, Chen H, Liu L, Mansfield KD, et al. Mitochondrial complex III is required for hypoxia-induced ROS production and cellular oxygen sensing. Cell Metab. 2005;1(6):401–8. doi: 10.1016/j.cmet.2005.05.001 [DOI] [PubMed] [Google Scholar]
  • 49.Fuhrmann DC, Brüne B. Mitochondrial composition and function under the control of hypoxia. Redox Biol. 2017;12:208–15. doi: 10.1016/j.redox.2017.02.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Krick N, Davies J, Coulson B, Sobrido-Cameán D, Miller M, Oswald MCW, et al. Heterogeneous responses to embryonic critical period perturbations within the Drosophila larval locomotor circuit. Cold Spring Harbor Laboratory. 2024. doi: 10.1101/2024.09.14.613036 [DOI] [Google Scholar]
  • 51.Günay C, Sieling FH, Dharmar L, Lin W-H, Wolfram V, Marley R, et al. Distal spike initiation zone location estimation by morphological simulation of ionic current filtering demonstrated in a novel model of an identified Drosophila motoneuron. PLoS Comput Biol. 2015;11(5):e1004189. doi: 10.1371/journal.pcbi.1004189 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Andersen SL. Trajectories of brain development: point of vulnerability or window of opportunity? Neurosci Biobehav Rev. 2003;27(1–2):3–18. doi: 10.1016/s0149-7634(03)00005-8 [DOI] [PubMed] [Google Scholar]
  • 53.Cicchetti D. Resilience under conditions of extreme stress: a multilevel perspective. World Psychiatry. 2010;9(3):145–54. doi: 10.1002/j.2051-5545.2010.tb00297.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Walton KD, Lieberman D, Llinás A, Begin M, Llinás RR. Identification of a critical period for motor development in neonatal rats. Neuroscience. 1992;51(4):763–7. doi: 10.1016/0306-4522(92)90517-6 [DOI] [PubMed] [Google Scholar]
  • 55.Moorman SJ, Cordova R, Davies SA. A critical period for functional vestibular development in zebrafish. Dev Dyn. 2002;223(2):285–91. doi: 10.1002/dvdy.10052 [DOI] [PubMed] [Google Scholar]
  • 56.Avitan L, Pujic Z, Mölter J, Van De Poll M, Sun B, Teng H, et al. Spontaneous activity in the zebrafish tectum reorganizes over development and is influenced by visual experience. Curr Biol. 2017;27(16):2407-2419.e4. doi: 10.1016/j.cub.2017.06.056 [DOI] [PubMed] [Google Scholar]
  • 57.Xie J, Jusuf PR, Bui BV, Goodbourn PT. Experience-dependent development of visual sensitivity in larval zebrafish. Sci Rep. 2019;9(1):18931. doi: 10.1038/s41598-019-54958-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Chin JSR, Phan T-AN, Albert LT, Keene AC, Duboué ER. Long lasting anxiety following early life stress is dependent on glucocorticoid signaling in zebrafish. Sci Rep. 2022;12(1):12826. doi: 10.1038/s41598-022-16257-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Hageter J, Starkey J, Horstick EJ. Thalamic regulation of a visual critical period and motor behavior. Cell Rep. 2023;42(4):112287. doi: 10.1016/j.celrep.2023.112287 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Kiral FR, Dutta SB, Linneweber GA, Hilgert S, Poppa C, Duch C, et al. Brain connectivity inversely scales with developmental temperature in Drosophila. Cell Rep. 2021;37(12):110145. doi: 10.1016/j.celrep.2021.110145 [DOI] [PubMed] [Google Scholar]
  • 61.Falibene A, Roces F, Rössler W, Groh C. Daily thermal fluctuations experienced by pupae via rhythmic nursing behavior increase numbers of mushroom body microglomeruli in the adult ant brain. Front Behav Neurosci. 2016;10:73. doi: 10.3389/fnbeh.2016.00073 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Groh C, Tautz J, Rössler W. Synaptic organization in the adult honey bee brain is influenced by brood-temperature control during pupal development. Proc Natl Acad Sci U S A. 2004;101(12):4268–73. doi: 10.1073/pnas.0400773101 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Wang G, Gordon TN, Rainwater S. Maximum voluntary temperature of insect larvae reveals differences in their thermal biology. J Thermal Biol. 2008;33(7):380–4. doi: 10.1016/j.jtherbio.2008.06.002 [DOI] [Google Scholar]
  • 64.Züfle P, Batista LL, Brandão SC, D’Uva G, Daniel C, Martelli C. Impact of developmental temperature on neural growth, connectivity, and function. Sci Adv. 2025;11(3):eadp9587. doi: 10.1126/sciadv.adp9587 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Crisp S, Evers JF, Fiala A, Bate M. The development of motor coordination in Drosophila embryos. Development. 2008;135(22):3707–17. doi: 10.1242/dev.026773 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Crisp SJ, Evers JF, Bate M. Endogenous patterns of activity are required for the maturation of a motor network. J Neurosci. 2011;31(29):10445–50. doi: 10.1523/JNEUROSCI.0346-11.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Coulson B, Hunter I, Doran S, Parkin J, Landgraf M, Baines RA. Critical periods in Drosophila neural network development: importance to network tuning and therapeutic potential. Front Physiol. 2022;13:1073307. doi: 10.3389/fphys.2022.1073307 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Graham C, Stefanatos R, Yek AEH, Spriggs RV, Loh SHY, Uribe AH, et al. Mitochondrial ROS signalling requires uninterrupted electron flow and is lost during ageing in flies. Geroscience. 2022;44(4):1961–74. doi: 10.1007/s11357-022-00555-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Weidenmüller A, Mayr C, Kleineidam CJ, Roces F. Preimaginal and adult experience modulates the thermal response behavior of ants. Curr Biol. 2009;19(22):1897–902. doi: 10.1016/j.cub.2009.08.059 [DOI] [PubMed] [Google Scholar]
  • 70.Becher MA, Scharpenberg H, Moritz RFA. Pupal developmental temperature and behavioral specialization of honeybee workers (Apis mellifera L.). J Comp Physiol A Neuroethol Sens Neural Behav Physiol. 2009;195(7):673–9. doi: 10.1007/s00359-009-0442-7 [DOI] [PubMed] [Google Scholar]
  • 71.Tautz J, Maier S, Groh C, Rossler W, Brockmann A. Behavioral performance in adult honey bees is influenced by the temperature experienced during their pupal development. Proc Natl Acad Sci U S A. 2003;100(12):7343–7. doi: 10.1073/pnas.1232346100 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Vicidomini R, Serpe M. Local BMP signaling: a sensor for synaptic activity that balances synapse growth and function. Curr Top Dev Biol. 2022;150:211–54. doi: 10.1016/bs.ctdb.2022.04.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Broadie KS, Bate M. Development of the embryonic neuromuscular synapse of Drosophila melanogaster. J Neurosci. 1993;13(1):144–66. doi: 10.1523/JNEUROSCI.13-01-00144.1993 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Baines RA, Bate M. Electrophysiological development of central neurons in the Drosophila embryo. J Neurosci. 1998;18(12):4673–83. doi: 10.1523/JNEUROSCI.18-12-04673.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Mòdol L, Moissidis M, Selten M, Oozeer F, Marín O. Somatostatin interneurons control the timing of developmental desynchronization in cortical networks. Neuron. 2024;112(12):2015-2030.e5. doi: 10.1016/j.neuron.2024.03.014 [DOI] [PubMed] [Google Scholar]
  • 76.Tien N-W, Kerschensteiner D. Homeostatic plasticity in neural development. Neural Dev. 2018;13(1):9. doi: 10.1186/s13064-018-0105-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Chen L, Li X, Tjia M, Thapliyal S. Homeostatic plasticity and excitation-inhibition balance: the good, the bad, and the ugly. Curr Opin Neurobiol. 2022;75:102553. doi: 10.1016/j.conb.2022.102553 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Swann JW, Rho JM. How is homeostatic plasticity important in epilepsy?. Adv Exp Med Biol. 2014;813:123–31. doi: 10.1007/978-94-017-8914-1_10 [DOI] [PubMed] [Google Scholar]
  • 79.Davis GW, Bezprozvanny I. Maintaining the stability of neural function: a homeostatic hypothesis. Annu Rev Physiol. 2001;63:847–69. doi: 10.1146/annurev.physiol.63.1.847 [DOI] [PubMed] [Google Scholar]
  • 80.O’Donovan MJ, Chub N, Wenner P. Mechanisms of spontaneous activity in developing spinal networks. J Neurobiol. 1998;37(1):131–45. doi: [DOI] [PubMed] [Google Scholar]
  • 81.Wong WT, Sanes JR, Wong RO. Developmentally regulated spontaneous activity in the embryonic chick retina. J Neurosci. 1998;18(21):8839–52. doi: 10.1523/JNEUROSCI.18-21-08839.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Dhawan S, Myers P, Bailey DMD, Ostrovsky AD, Evers JF, Landgraf M. Reactive oxygen species mediate activity-regulated dendritic plasticity through NADPH oxidase and aquaporin regulation. Front Cell Neurosci. 2021;15:641802. doi: 10.3389/fncel.2021.641802 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Robb EL, Hall AR, Prime TA, Eaton S, Szibor M, Viscomi C, et al. Control of mitochondrial superoxide production by reverse electron transport at complex I. J Biol Chem. 2018;293(25):9869–79. doi: 10.1074/jbc.RA118.003647 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Truszkowski TLS, Aizenman CD. Neurobiology: setting the set point for neural homeostasis. Curr Biol. 2015;25(23):R1132-3. doi: 10.1016/j.cub.2015.10.021 [DOI] [PubMed] [Google Scholar]
  • 85.Styr B, Gonen N, Zarhin D, Ruggiero A, Atsmon R, Gazit N, et al. Mitochondrial regulation of the hippocampal firing rate set point and seizure susceptibility. Neuron. 2019;102(5):1009-1024.e8. doi: 10.1016/j.neuron.2019.03.045 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Vander Heiden MG, Cantley LC, Thompson CB. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science. 2009;324(5930):1029–33. doi: 10.1126/science.1160809 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Wang C-W, Purkayastha A, Jones KT, Thaker SK, Banerjee U. In vivo genetic dissection of tumor growth and the Warburg effect. Elife. 2016;5:e18126. doi: 10.7554/eLife.18126 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Elzakra N, Kim Y. HIF-1α metabolic pathways in human cancer. Adv Exp Med Biol. 2021;1280:243–60. doi: 10.1007/978-3-030-51652-9_17 [DOI] [PubMed] [Google Scholar]
  • 89.Dunwoodie SL. The role of hypoxia in development of the Mammalian embryo. Dev Cell. 2009;17(6):755–73. doi: 10.1016/j.devcel.2009.11.008 [DOI] [PubMed] [Google Scholar]
  • 90.Bohuslavova R, Cerychova R, Papousek F, Olejnickova V, Bartos M, Görlach A, et al. HIF-1α is required for development of the sympathetic nervous system. Proc Natl Acad Sci U S A. 2019;116(27):13414–23. doi: 10.1073/pnas.1903510116 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Kallio PJ, Okamoto K, O’Brien S, Carrero P, Makino Y, Tanaka H, et al. Signal transduction in hypoxic cells: inducible nuclear translocation and recruitment of the CBP/p300 coactivator by the hypoxia-inducible factor-1alpha. EMBO J. 1998;17(22):6573–86. doi: 10.1093/emboj/17.22.6573 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Perez-Perri JI, Dengler VL, Audetat KA, Pandey A, Bonner EA, Urh M, et al. The TIP60 complex is a conserved coactivator of HIF1A. Cell Rep. 2016;16(1):37–47. doi: 10.1016/j.celrep.2016.05.082 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Luo W, Wang Y. Epigenetic regulators: multifunctional proteins modulating hypoxia-inducible factor-α protein stability and activity. Cell Mol Life Sci. 2018;75(6):1043–56. doi: 10.1007/s00018-017-2684-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Yfantis A, Mylonis I, Chachami G, Nikolaidis M, Amoutzias GD, Paraskeva E, et al. Transcriptional response to hypoxia: the role of HIF-1-associated co-regulators. Cells. 2023;12(5):798. doi: 10.3390/cells12050798 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Fujioka M, Lear BC, Landgraf M, Yusibova GL, Zhou J, Riley KM, et al. Even-skipped, acting as a repressor, regulates axonal projections in Drosophila. Development. 2003;130(22):5385–400. doi: 10.1242/dev.00770 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Han Z, Fujioka M, Su M, Liu M, Jaynes JB, Bodmer R. Transcriptional integration of competence modulated by mutual repression generates cell-type specificity within the cardiogenic mesoderm. Dev Biol. 2002;252(2):225–40. doi: 10.1006/dbio.2002.0846 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Radyuk SN, Rebrin I, Klichko VI, Sohal BH, Michalak K, Benes J, et al. Mitochondrial peroxiredoxins are critical for the maintenance of redox state and the survival of adult Drosophila. Free Radic Biol Med. 2010;49(12):1892–902. doi: 10.1016/j.freeradbiomed.2010.09.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Sulkowski MJ, Han TH, Ott C, Wang Q, Verheyen EM, Lippincott-Schwartz J, et al. A novel, noncanonical BMP pathway modulates synapse maturation at the Drosophila neuromuscular junction. PLoS Genet. 2016;12(1):e1005810. doi: 10.1371/journal.pgen.1005810 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Laker RC, Xu P, Ryall KA, Sujkowski A, Kenwood BM, Chain KH, et al. A novel MitoTimer reporter gene for mitochondrial content, structure, stress, and damage in vivo. J Biol Chem. 2014;289(17):12005–15. doi: 10.1074/jbc.M113.530527 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Bischof J, Maeda RK, Hediger M, Karch F, Basler K. An optimized transgenesis system for Drosophila using germ-line-specific phiC31 integrases. Proc Natl Acad Sci U S A. 2007;104(9):3312–7. doi: 10.1073/pnas.0611511104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Baines RA, Bate M. Electrophysiological development of central neurons in the Drosophila embryo. J Neurosci. 1998;18(12):4673–83. doi: 10.1523/JNEUROSCI.18-12-04673.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Risse B, Berh D, Otto N, Klämbt C, Jiang X. FIMTrack: an open source tracking and locomotion analysis software for small animals. PLoS Comput Biol. 2017;13(5):e1005530. doi: 10.1371/journal.pcbi.1005530 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Risse B, Thomas S, Otto N, Löpmeier T, Valkov D, Jiang X, et al. FIM, a novel FTIR-based imaging method for high throughput locomotion analysis. PLoS One. 2013;8(1):e53963. doi: 10.1371/journal.pone.0053963 [DOI] [PMC free article] [PubMed] [Google Scholar]

Decision Letter 0

Christian Schnell, PhD

3 Feb 2025

Dear Dr Sobrid-Camean,

Thank you for submitting your manuscript entitled "Mitochondrial ROS and HIF-1α signalling mediate critical period plasticity" for consideration as a Research Article by PLOS Biology.

Your manuscript has now been evaluated by the PLOS Biology editorial staff as well as by an academic editor with relevant expertise and I am writing to let you know that we would like to send your submission out for external peer review.

However, before we can send your manuscript to reviewers, we need you to complete your submission by providing the metadata that is required for full assessment. To this end, please login to Editorial Manager where you will find the paper in the 'Submissions Needing Revisions' folder on your homepage. Please click 'Revise Submission' from the Action Links and complete all additional questions in the submission questionnaire.

Once your full submission is complete, your paper will undergo a series of checks in preparation for peer review. After your manuscript has passed the checks it will be sent out for review. To provide the metadata for your submission, please Login to Editorial Manager (https://www.editorialmanager.com/pbiology) within two working days, i.e. by Feb 05 2025 11:59PM.

If your manuscript has been previously peer-reviewed at another journal, PLOS Biology is willing to work with those reviews in order to avoid re-starting the process. Submission of the previous reviews is entirely optional and our ability to use them effectively will depend on the willingness of the previous journal to confirm the content of the reports and share the reviewer identities. Please note that we reserve the right to invite additional reviewers if we consider that additional/independent reviewers are needed, although we aim to avoid this as far as possible. In our experience, working with previous reviews does save time.

If you would like us to consider previous reviewer reports, please edit your cover letter to let us know and include the name of the journal where the work was previously considered and the manuscript ID it was given. In addition, please upload a response to the reviews as a 'Prior Peer Review' file type, which should include the reports in full and a point-by-point reply detailing how you have or plan to address the reviewers' concerns.

During the process of completing your manuscript submission, you will be invited to opt-in to posting your pre-review manuscript as a bioRxiv preprint. Visit http://journals.plos.org/plosbiology/s/preprints for full details. If you consent to posting your current manuscript as a preprint, please upload a single Preprint PDF.

Feel free to email us at plosbiology@plos.org if you have any queries relating to your submission.

Kind regards,

Christian

Christian Schnell, PhD

Senior Editor

PLOS Biology

cschnell@plos.org

Decision Letter 1

Christian Schnell, PhD

12 Mar 2025

Dear Dr Sobrid-Camean,

Thank you for your patience while your manuscript "Mitochondrial ROS and HIF-1α signalling mediate critical period plasticity" was peer-reviewed at PLOS Biology. It has now been evaluated by the PLOS Biology editors, an Academic Editor with relevant expertise, and by several independent reviewers.

In light of the reviews, which you will find at the end of this email, we would like to invite you to revise the work to thoroughly address the reviewers' reports.

As you will see below, the reviewers agree that the study is overall well done and provides important insights. Reviewer 2 has only a few minor suggestions, but Reviewer 1 and Reviewer 3 request a number of additional control experiments and analyses that will be required to fully address their concerns.

Given the extent of revision needed, we cannot make a decision about publication until we have seen the revised manuscript and your response to the reviewers' comments. Your revised manuscript is likely to be sent for further evaluation by all or a subset of the reviewers.

In addition to these revisions, you will need to complete some formatting changes, which you will receive in a follow up email. A member of our team will be in touch with a set of requests shortly.

We expect to receive your revised manuscript within 3 months. Please email us (plosbiology@plos.org) if you have any questions or concerns, or would like to request an extension.

At this stage, your manuscript remains formally under active consideration at our journal; please notify us by email if you do not intend to submit a revision so that we may withdraw it.

**IMPORTANT - SUBMITTING YOUR REVISION**

Your revisions should address the specific points made by each reviewer. Please submit the following files along with your revised manuscript:

1. A 'Response to Reviewers' file - this should detail your responses to the editorial requests, present a point-by-point response to all of the reviewers' comments, and indicate the changes made to the manuscript.

*NOTE: In your point-by-point response to the reviewers, please provide the full context of each review. Do not selectively quote paragraphs or sentences to reply to. The entire set of reviewer comments should be present in full and each specific point should be responded to individually, point by point.

You should also cite any additional relevant literature that has been published since the original submission and mention any additional citations in your response.

2. In addition to a clean copy of the manuscript, please also upload a 'track-changes' version of your manuscript that specifies the edits made. This should be uploaded as a "Revised Article with Changes Highlighted" file type.

*Re-submission Checklist*

When you are ready to resubmit your revised manuscript, please refer to this re-submission checklist: https://plos.io/Biology_Checklist

To submit a revised version of your manuscript, please go to https://www.editorialmanager.com/pbiology/ and log in as an Author. Click the link labelled 'Submissions Needing Revision' where you will find your submission record.

Please make sure to read the following important policies and guidelines while preparing your revision:

*Published Peer Review*

Please note while forming your response, if your article is accepted, you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out. Please see here for more details:

https://blogs.plos.org/plos/2019/05/plos-journals-now-open-for-published-peer-review/

*PLOS Data Policy*

Please note that as a condition of publication PLOS' data policy (http://journals.plos.org/plosbiology/s/data-availability) requires that you make available all data used to draw the conclusions arrived at in your manuscript. If you have not already done so, you must include any data used in your manuscript either in appropriate repositories, within the body of the manuscript, or as supporting information (N.B. this includes any numerical values that were used to generate graphs, histograms etc.). For an example see here: http://www.plosbiology.org/article/info%3Adoi%2F10.1371%2Fjournal.pbio.1001908#s5

*Blot and Gel Data Policy*

We require the original, uncropped and minimally adjusted images supporting all blot and gel results reported in an article's figures or Supporting Information files. We will require these files before a manuscript can be accepted so please prepare them now, if you have not already uploaded them. Please carefully read our guidelines for how to prepare and upload this data: https://journals.plos.org/plosbiology/s/figures#loc-blot-and-gel-reporting-requirements

*Protocols deposition*

To enhance the reproducibility of your results, we recommend that if applicable you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols

Thank you again for your submission to our journal. We hope that our editorial process has been constructive thus far, and we welcome your feedback at any time. Please don't hesitate to contact us if you have any questions or comments.

Sincerely,

Christian

Christian Schnell, PhD

Senior Editor

PLOS Biology

cschnell@plos.org

------------------------------------

REVIEWS:

Reviewer #1 (James Jepson): Mitochondrial ROS and HIF-1α signalling mediate critical period plasticity. Sobrido-Cameán et al., (2025).

Overview:

In this work, Sobrido-Cameán and colleagues investigate the molecular mechanisms underlying a critical period of sensitivity to heat stress during the development of the Drosophila larval locomotor network.

The authors show that heat stress during a narrow window of developmental time (17-19 h after egg laying) causes long-lasting alterations in pre- and post-synaptic development and neural excitability during the larval stages of the Drosophila lifecycle. Using elegant genetic approaches, they identify mitochondrial reactive oxygen species (ROS) as a key mediator of these alterations. Furthermore, they provide evidence that the conserved epigenetic regulator HIF-1alpha/Sima acts downstream of ROS to drive these long-term changes in development and intrinsic excitability.

The identification of ROS and HIF-1alpha as effectors of a critical period influencing locomotion is both novel and exciting, and may help further elucidate the molecular underpinnings of critical periods during neural development. Since these developmental phases are disrupted in a range of neurological disorders, this work both advances the fundamental understanding of neural development and has potential clinical relevance.

The studies themselves are well performed and include technically challenging electrophysiological recordings from larval motoneurons. There are certain controls that could be added to improve robustness. Additional analyses of existing data and textual alterations could provide further mechanistic insights and help the readers understand the work more clearly. Finally, it would be highly beneficial if the authors could directly test for increases in pre- and post-synaptic ROS in response to heat stress, as predicted by their model and prior data. With these inclusions, I would be very happy to see this interesting work published in PLoS Biology.

Major comments:

1. ROS levels.

The authors use several independent genetic manipulations to show that ROS is required post-synaptically for heat-induced changes in post-synaptic GluR11A levels, and presynaptic bouton number (Figure 1), pre-synaptically for heat-induced changes in presynaptic bouton number (Figure 3), and pre-synaptically for heat-induced changes in intrinsic motoneuron excitability (Figure 5). The authors mention prior work showing that ROS can mediate activity-dependent synaptic growth (a presynaptic effect), and that heat stress elevates ROS (page 3). However, they do not directly show that in their experimental paradigm, embryonic heat stress increases ROS, and importantly, whether this occur both pre- and post-synaptically (as predicted).

I agree with the authors that this is likely to occur, but if it did not, it would significantly alter their mechanistic model. Therefore, it would be beneficial for the authors to test whether ROS is indeed elevated by heat stress during the embryonic critical period, in motoneurons and the post-synaptic muscle tissue, perhaps using transgenic reporters such as UAS-mito-roGFP2-Orp1, as they have used previously (Oswald et al., (2018) eLife).

2. Methods for transient transgene expression.

The schematics in Figures 1-6 depict transient control of transgene expression during (approximately) the embryonic critical period. How this is achieved is not immediately clear to the reader - the AGES system is only described after Fig. 3B-D are mentioned (for Supp. Fig. 1), and in the legend of Fig. 5. Can the authors please clearly explain to the reader how they control transgene activity at the earliest stage (i.e for Figure 1), and mention the logic of using the AGES system? (I presume this was used because the more commonly used TARGET system requires temperature changes?).

This leads to an important control condition that appears to be currently missing in some datasets - an auxin-fed transgene or driver-alone control containing auxin-Gal80. This genotype appears present in Figure 5 and Supp. Figs 1 and 2, but it was not clear to me whether this control was also fed auxin at the same time as the experimental genotypes (apologies if I missed it). Please clarify in the text and add this control to relevant datasets if absent.

3. RNAi line controls.

The authors use various RNAi lines (#26207, 50727, 34717) that are in a y, sc, v, sev quadruple mutant background (Figures 4-5). Have the authors tested RNAi alone controls to confirm that synaptic development is not altered in male hemizygotes or heterozygote females for these mutants?

4. Morphological analyses at the NMJ.

At the larval NMJ, the authors focus on the effect of heat stress on presynaptic bouton number and post-synaptic GluR expression. The images provided by the authors generated questions to me about two other aspects of NMJ development. Firstly, whether heat stress also affects bouton size? This seems to decrease somewhat in the 32C control, but unlike bouton number, this property does not seem altered by AOX over-expression (at least, in the image provided; Fig. 1B). This raises the question of whether some morphological alterations are driven by ROS, but others are ROS-independent? Secondly, have the authors quantified ghost boutons at the NMJ (PMID: 18341991), and tested whether their number is affected by heat stress? Again, images in Fig. 1B are suggestive of potential alterations. A decrease in ghost bouton number might suggest a pathway to increase mature bouton number.

These suggestions are not essential to address but could yield interesting mechanistic insights.

5. Presynaptic and postsynaptic effects of ROS/HIF-1alpha.

The authors identify a mix of pre- and post-synaptic requirements for ROS and HIF-1alpha in driving heat-induced changes in development during the critical period. ROS are required post-synaptically for changes in bouton number and GluRIIA levels Fig.1), and pre-synaptically for increased bouton and active zone number (Fig.3) and reduced intrinsic excitability (Fig.5); while HIF1-alpha is required post-synaptically for changes in bouton number and GluRIIA levels (Fig.4), and pre-synaptically for reduced intrinsic excitability (Fig.5).

These complex requirements emphasize the importance of determining where ROS is increasing (see above) in response to heat stress. The results also suggest that post-synaptic ROS/HIF1-alpha induce retrograde signals that impact presynaptic bouton number. However, it is unclear whether the reverse is true: can enhancing presynaptic ROS/HIF1-alpha act trans-synaptically to alter post-synaptic GluRIIA levels?

In the context of temperature-induced changes in larval crawling, the above results also made we wonder whether the authors had tested how muscle-specific manipulations of AOX/HIF1 affected crawling speed; and/or whether the authors had distinguished whether this pathway was relevant solely in motoneurons (using a motoneuron-specific driver) for adaptations to heat stress, or were required more broadly throughout the CNS?

Again, these are not essential experiments but their inclusion might help more narrowly define relevant domains in which the ROS/HIF1 axis is important during the critical period.

6. Basis of reduced intrinsic excitability (Figure 5).

The authors show that heat stress during the critical period reduced AP firing rate without altering RMP. How do the authors stipulate that AP frequency is being altered? Could the authors examine AP shapes in their datasets to test whether this property is affected by embryonic heat stress? For example, could the duration/amplitude of the afterhyperpolarization current be increased? Or spike width?

Minor comments/questions:

1. How do the authors know that feeding auxin to gravid females induces transgene expression during the latter half of the embryonic stage? (Supp. Fig. 1A).

2. Descriptions of manipulations. E.g Fig.5, p.6. For all figures, can the authors please specific in the text whether transgene manipulation is occurring pre- or post-synaptically? i.e the effect is cell-autonomous or non-cell-autonomous.

3. Figure 5: it was unclear to me why HIF-1alpha and VHL manipulations are combined here (Fig.5C-D) but not elsewhere in the text. Can the authors please clarify and provide rationale?

4. Please define the 'CNS-Gal4' - I presume this is R57C10-Gal4 (Table 1), but I could not find a statement to this effect.

5. Discussion: the authors state they 'demonstrate the cell-autonomous necessity of RET-generated mitochondrial ROS and HIF-1a signalling for heat stress-induced critical period plasticity' (p.8). However, many effects described in this manuscript are non-cell-autonomous (e.g retrograde changes at the NMJ). Can the authors clarify that some of the effects are cell-autonomous, whereas others are non-cell-autonomous?

Reviewer #2: Sobrido-Cameán et al. demonstrate that temperature stress during the late embryonic critical period (CP) induces synaptic plasticity at the Drosophila neuromuscular junction. Using mutant lines and analyses of presynaptic morphology, glutamate receptor expression, synaptic active zones, electrophysiology, and crawling behavior, they reveal that CP plasticity is driven by ROS generation via reverse electron transfer, leading to HIF1-alpha induction.

Comments:

- The legend of figure 4 is missing and seems to be copy-pasted from figure 2.

- A recent study (PMID: 39721493) showed onset of presynaptic plasticity after acute emission of mitochondrial hydrogen peroxide in late-stage larvae. The authors should cite this work and discuss how it relates to their finding that RET-ROS induces plasticity if induced during but not after the critical period (supp fig 1).

Reviewer #3: The importance of critical periods during neurodevelopment cannot be understated; disruption of synaptic activity during developmental critical periods has lasting effects on synaptic morphology, growth, function, and behavioral coordination. Though the existence of critical periods has long been known, the events that occur during those critical periods and importantly, the molecules needed to enact them have remained considerably more mysterious. Sobrido-Cameán and colleagues here identify two fundamental new concepts regarding critical period biology: first, that manipulating a single cell during a critical period can influence later changes in morphology, protein recruitment, and activity; and second, that reactive oxygen species (ROS) from mitochondria and acting through HIF-1α are required to mediate critical period signals. Mitochondrial ROS is required to signal in postsynaptic muscles or presynaptic motoneurons - in the absence of ROS (either through failure to generate or subsequent degradation), critical period changes induced by high temperature bouts are blocked. These changes are seen morphologically (involving bouton number, GluRIIA abundance, or active zone number) and functionally (involving action potential number or crawling ability. In muscles, blocking (either by directly impairing or expressing enzymes that regulate its levels) HIF-1α precludes ability to induce changes during the critical period.

The findings represent a significant advance in our understanding of the molecular basis of critical period signaling. Moreover, the data posit a fascinating new addition to the burgeoning concept that ROS are not just a deleterious byproduct of existence, but rather, an important neuronal signaling axis. The experiments are particularly well designed and the genetics very well controlled, with examples of loss-of-function and gain-of-function consistent with each other. Moreover, the use of exogenously applied constructs (like yeast Ndi1 and AOX) to demonstrate similar manipulations of ROS is particularly helpful. There are, however, a few points that should be addressed in a revised manuscript.

Major Issues:

1) Does this work transsynaptically? When mitochondrial ROS is manipulated in motoneurons, does this influence GluRIIA abundance? When mitochondrial ROS or HIF-1α is manipulated in muscle, does this influence active zones? Similarly (if possible), does concurrent transsynaptic manipulation enhance the effects during the critical period or does one cell basically hit the ceiling of how much the system can change?

2) The main figures focus on HIF-1α in muscles with some evidence from the supplement that HIF-1α may also function in motoneurons? Can this data be elevated from the supplement to the main paper and also expanded upon? It is helpful to know if the same downstream effector occurs pre- and postsynaptically.

3) Specifically regarding the change in active zone number: do ROS manipulations alter active zone density? The changes show an increase in bouton number, so naturally, it would be expected if active zone number scales with bouton number (as has been shown developmentally), then this would naturally occur. Therefore, this is expected - unless there is some shift in active zone density, which should be examined.

4) When Ndi1 is expressed in DA1 at 25 degrees, there is an induction of synaptic changes during the critical period? If this manipulation is done at 32 degrees, do the two enhance each other? Or do they appear similar to a single manipulation? This would be helpful in ensuring that (at least genetically), they function in the same oeuvre.

5) The authors use multiple methods to regulate HIF-1α levels. Are there suitable reagents to examine HIF-1a levels to validate these methods in cells examined?

Minor Issues:

1) Though I understand the concept of Figure 5B, it would be very challenging for a non-physiologist to see the phenotype the authors are describing? Could you consider a different way of presenting? Perhaps picking one specific current and showing the results? Are the changes significant? Do they trend?

2) Though Table 1 is helpful, strain source data should be in the first Materials and Methods section. It would be much easier that way to determine provenance of the reagents used.

Decision Letter 2

Christian Schnell, PhD

21 Jul 2025

Dear Daniel,

Thank you for your patience while we considered your revised manuscript "Mitochondrial ROS and HIF-1α signalling mediate critical period plasticity" for publication as a Research Article at PLOS Biology. This revised version of your manuscript has been evaluated by the PLOS Biology editors, the Academic Editor and two of the original reviewers.

Based on the reviews and on our Academic Editor's assessment of your revision, we are likely to accept this manuscript for publication, provided you satisfactorily address the remaining points raised by the reviewers. Please also make sure to address the following data and other policy-related requests:

* We would like to suggest a different title to improve its accessibility for our broad audience: "Mitochondrial ROS and HIF-1α signaling mediate synaptic plasticity in the critical period"

* Please add the links to the funding agencies in the Financial Disclosure statement in the manuscript details.

* DATA POLICY:

You may be aware of the PLOS Data Policy, which requires that all data be made available without restriction: http://journals.plos.org/plosbiology/s/data-availability. For more information, please also see this editorial: http://dx.doi.org/10.1371/journal.pbio.1001797

Note that we do not require all raw data. Rather, we ask that all individual quantitative observations that underlie the data summarized in the figures and results of your paper be made available in one of the following forms:

1) Supplementary files (e.g., excel). Please ensure that all data files are uploaded as 'Supporting Information' and are invariably referred to (in the manuscript, figure legends, and the Description field when uploading your files) using the following format verbatim: S1 Data, S2 Data, etc. Multiple panels of a single or even several figures can be included as multiple sheets in one excel file that is saved using exactly the following convention: S1_Data.xlsx (using an underscore).

2) Deposition in a publicly available repository. Please also provide the accession code or a reviewer link so that we may view your data before publication.

Regardless of the method selected, please ensure that you provide the individual numerical values that underlie the summary data displayed in the following figure panels as they are essential for readers to assess your analysis and to reproduce it: 1BDEF, 2CDE, 3BDEFG, 4DE, 5DEGHIJK, S1AB, S2BD, S4CD, S5B and S6AB.

NOTE: the numerical data provided should include all replicates AND the way in which the plotted mean and errors were derived (it should not present only the mean/average values).

Please also ensure that figure legends in your manuscript include information on where the underlying data can be found, and ensure your supplemental data file/s has a legend.

Please ensure that your Data Statement in the submission system accurately describes where your data can be found.

* CODE POLICY

Per journal policy, if you have generated any custom code during the course of this investigation, please make it available without restrictions. Please ensure that the code is sufficiently well documented and reusable, and that your Data Statement in the Editorial Manager submission system accurately describes where your code can be found.

Please note that we cannot accept sole deposition of code in GitHub, as this could be changed after publication. However, you can archive this version of your publicly available GitHub code to Zenodo. Once you do this, it will generate a DOI number, which you will need to provide in the Data Accessibility Statement (you are welcome to also provide the GitHub access information). See the process for doing this here: https://docs.github.com/en/repositories/archiving-a-github-repository/referencing-and-citing-content

As you address these items, please take this last chance to review your reference list to ensure that it is complete and correct. If you have cited papers that have been retracted, please include the rationale for doing so in the manuscript text, or remove these references and replace them with relevant current references. Any changes to the reference list should be mentioned in the cover letter that accompanies your revised manuscript.

In addition to these revisions, you may need to complete some formatting changes, which you will receive in a follow up email. A member of our team will be in touch with a set of requests shortly. If you do not receive a separate email within a few days, please assume that checks have been completed, and no additional changes are required.

We expect to receive your revised manuscript within two weeks.

To submit your revision, please go to https://www.editorialmanager.com/pbiology/ and log in as an Author. Click the link labelled 'Submissions Needing Revision' to find your submission record. Your revised submission must include the following:

- a cover letter that should detail your responses to any editorial requests, if applicable, and whether changes have been made to the reference list

- a Response to Reviewers file that provides a detailed response to the reviewers' comments (if applicable, if not applicable please do not delete your existing 'Response to Reviewers' file.)

- a track-changes file indicating any changes that you have made to the manuscript.

NOTE: If Supporting Information files are included with your article, note that these are not copyedited and will be published as they are submitted. Please ensure that these files are legible and of high quality (at least 300 dpi) in an easily accessible file format. For this reason, please be aware that any references listed in an SI file will not be indexed. For more information, see our Supporting Information guidelines:

https://journals.plos.org/plosbiology/s/supporting-information

*Published Peer Review History*

Please note that you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out. Please see here for more details:

https://plos.org/published-peer-review-history/

*Press*

Should you, your institution's press office or the journal office choose to press release your paper, please ensure you have opted out of Early Article Posting on the submission form. We ask that you notify us as soon as possible if you or your institution is planning to press release the article.

*Protocols deposition*

To enhance the reproducibility of your results, we recommend that if applicable you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols

Please do not hesitate to contact me should you have any questions.

Sincerely,

Christian

Christian Schnell, PhD

Senior Editor

cschnell@plos.org

PLOS Biology

------------------------------------------------------------------------

Reviewer remarks:

Reviewer #1: The authors have fully addressed my initial comments. In their revised text, they have included both new data and further analyses of thier original datasets that have collectively enhanced the robustness of their findings and conclusions. Of particular note, they have generated and utilised a novel transgenic reporter of ROS to demonstrate that increasing ambient temperature during the embryonic critical period indeed enhances ROS levels both pre- and post-synaptically - a finding that adds support to their mechanistic model. They have also modified the text to further enhance the clarity of the manuscript. I congratulate the authors on their excellent work, which has revealed exciting new insights into the molecular basis of critical periods influencing motor circuit development. I look forward to seeing their manuscript in press.

Reviewer #3: The authors have done a commendable job addressing reviewer comments and I applaud them for the careful thought, rigorous consideration, and the additional new data provided. I think this is suitable for publication save one minor adjustment:

1) I appreciate the addition of DA1 muscle and aCC neuron diagrams but their current iteration is far too small for them to have any true added benefit to the figures. Can these diagrams be made much bigger so that the reader can understand better the types of experiments being done?

Decision Letter 3

Christian Schnell, PhD

30 Jul 2025

Dear Daniel,

Thank you for the submission of your revised Research Article "Mitochondrial ROS and HIF-1α signaling mediate synaptic plasticity in the critical period" for publication in PLOS Biology. On behalf of my colleagues and the Academic Editor, Timothy Mosca, I am pleased to say that we can in principle accept your manuscript for publication, provided you address any remaining formatting and reporting issues. These will be detailed in an email you should receive within 2-3 business days from our colleagues in the journal operations team; no action is required from you until then. Please note that we will not be able to formally accept your manuscript and schedule it for publication until you have completed any requested changes.

Please take a minute to log into Editorial Manager at http://www.editorialmanager.com/pbiology/, click the "Update My Information" link at the top of the page, and update your user information to ensure an efficient production process.

PRESS

We frequently collaborate with press offices. If your institution or institutions have a press office, please notify them about your upcoming paper at this point, to enable them to help maximise its impact. If the press office is planning to promote your findings, we would be grateful if they could coordinate with biologypress@plos.org. If you have previously opted in to the early version process, we ask that you notify us immediately of any press plans so that we may opt out on your behalf.

We also ask that you take this opportunity to read our Embargo Policy regarding the discussion, promotion and media coverage of work that is yet to be published by PLOS. As your manuscript is not yet published, it is bound by the conditions of our Embargo Policy. Please be aware that this policy is in place both to ensure that any press coverage of your article is fully substantiated and to provide a direct link between such coverage and the published work. For full details of our Embargo Policy, please visit http://www.plos.org/about/media-inquiries/embargo-policy/.

Thank you again for choosing PLOS Biology for publication and supporting Open Access publishing. We look forward to publishing your study. 

Sincerely, 

Christian

Christian Schnell, PhD

Senior Editor

PLOS Biology

cschnell@plos.org

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Supplementary Materials

    S1 Fig. Control of UAS-lines crossed to wild type Oregon-R.

    (A) Dot-plot quantification shows changes to aCC NMJ growth on its target muscle DA1, based on the standard measure of the number of boutons (swellings containing multiple presynaptic release sites/active zones). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the wild type of condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with wild type exposed to 32°C heat stress during the embryonic critical period. “Wild type” is Oregon-R. (B) Dot-plot quantification shows changes in levels of the GluRIIA receptor subunit at aCC NMJs quantified in C). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the wild type condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with wild type exposed to 32°C heat stress during the embryonic critical period. “Wild type” is Oregon-R. See raw data in S6 Data.

    (TIFF)

    pbio.3003338.s001.tiff (419.5KB, tiff)
    S2 Fig. Muscles influence motoneuron at NMJ during the critical period, but motoneuron do not affect GluR composition in muscles.

    (A) Heat stress experienced during the embryonic critical period (32°C vs. 25°C control) leads to decreased postsynaptic GluRIIA. Simultaneous genetic manipulation of aCC motoneuron during embryonic stages does not affect GluRIIA in muscles. “Control” indicates control genotype heterozygous for Oregon-R and aCC-GAL4[1]. Larvae were reared at the control temperature of 25°C until the late wandering stage, 100 h after larval hatching (ALH). Data are shown with mean ± SEM, ANOVA, **p < 0.001, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. (B) Heat stress experienced during the embryonic critical period (32°C vs. 25°C control) leads to increased presynaptic active zone number. Simultaneous genetic manipulation of muscle DA1 during embryonic stages suggests that ROS by Complex-I is necessary and sufficient to induce active zones changes presynaptically. “Control” indicates control genotype heterozygous for Oregon-R and DA1-GAL4. Larvae were reared at the control temperature of 25°C until the late wandering stage, 100 h after larval hatching (ALH). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. Black asterisks indicate comparison with the control condition of 25°C throughout, genetically unmanipulated. Red asterisks indicate comparisons with control genotype exposed to 32°C heat stress during the embryonic critical period. See raw data in S7 Data.

    (TIFF)

    pbio.3003338.s002.tiff (478.6KB, tiff)
    S3 Fig. Characterization of aCC-GAL4 [2] expression.

    Expression starts around 13 h after egg lay (AEL) in few sporadic cells, not including the aCC motoneuron and mostly limited to the head region. First aCC motoneurons show GAL4 expression as of 14.5 h AEL, with more aCC motoneurons expressing from 15 h AEL onwards (i.e., 2 h prior to critical period opening). Concomitantly, GAL4 expression in other cells disappears. From laraval hatching onwards, segmentally repeated expression in all aCC motoneurons is maintained until at least 72 h after larval hatching (mid-3rd instar stage). Scale bar = 100 µm.

    (TIFF)

    pbio.3003338.s003.tiff (898.2KB, tiff)
    S4 Fig. RET-ROS and HIF-1α signaling has lasting impact on NMJ only during the critical period.

    (A) Experimental paradigm. (B) Genetic manipulation of motoneuron aCC during the critical period versus after critical period closure. “Control” indicates control genotype heterozygous for Oregon-R and aCC-GAL4[2]; Auxin-GAL80. Larvae were reared at the control temperature of 25oC until the late wandering stage, 100 h ALH. Scale bar = 20 µm. (C) Dot-plot quantification shows changes to aCC NMJ growth on its target muscle DA1, based on the standard measure of the number of boutons (swellings containing multiple presynaptic release sites/active zones). Data are shown with mean ± SEM, ANOVA, ****p < 0.00001, ‘ns’ indicates statistical non-significance. (D) Dot-plot quantification shows changes in the number of active zones at aCC NMJs quantified in C). Data are shown with mean ± SEM, ANOVA, ***p < 0.0001, ****p < 0.00001, ‘ns’ indicates statistical non-significance. See raw data in S8 Data.

    (TIFF)

    pbio.3003338.s004.tiff (775.3KB, tiff)
    S5 Fig. RET-ROS and HIF-1α signaling have no lasting impact on behavior after the critical period.

    (A) Experimental paradigm. (B) Crawling speed of third instar larvae (72 h after larval hatching (ALH)). Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress) and simultaneous genetic manipulation of all neurons during first instar stage. “Control” indicates control genotype heterozygous for Oregon-R and CNS-GAL4; Auxin-GAL80. Larvae were reared at the control temperature of 25°C until 72 h ALH. Each data point represents crawling speed from an individual uninterrupted continuous forward crawl, n = specimen replicate number, up to three crawls assayed for each larva. Mean ± SEM, ANOVA, ns = not significant, ****p < 0.00001. See raw in S9 Data.

    (TIFF)

    pbio.3003338.s005.tiff (303.3KB, tiff)
    S6 Fig. Homeostatic responses to environmental change are maintained after 32°C critical period.

    (A) Acute shift of temperature. Crawling speed of third instar larvae (72 h after larval hatching (ALH)). Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress); temperature experienced during the larva rearing (25oC control vs. 32°C heat stress) and control (25°C) versus acute shift of temperature during the crawling assay (32oC). Genotype is Oregon-R. Each data point represents crawling speed from an individual uninterrupted continuous forward crawl, n = specimen replicate number, up to three crawls assayed for each larva. Mean ± SEM, ANOVA, ns = not significant, ****p < 0.00001. (B) Chronic adaptation to the heat stress. Crawling speed of third instar larvae (72 h after larval hatching (ALH)). Temperature experienced during the embryonic critical period (25°C control vs. 32°C heat stress); temperature experienced during the larva rearing and during the crawling assay (25°C control vs. 32°C chronic adaptation). Oregon-R. Each data point represents crawling speed from an individual uninterrupted continuous forward crawl, n = specimen replicate number, up to three crawls assayed for each larva. Mean ± SEM, ANOVA, ns = not significant. See raw data in S10 Data.

    (TIFF)

    pbio.3003338.s006.tiff (336.4KB, tiff)
    S1 Data. Raw date of results in Fig 1.

    (XLSX)

    pbio.3003338.s007.xlsx (23.1KB, xlsx)
    S2 Data. Raw date of results in Fig 2.

    (XLSX)

    pbio.3003338.s008.xlsx (14.1KB, xlsx)
    S3 Data. Raw date of results in Fig 3.

    (XLSX)

    pbio.3003338.s009.xlsx (19.3KB, xlsx)
    S4 Data. Raw date of results in Fig 4.

    (XLSX)

    pbio.3003338.s010.xlsx (16.5KB, xlsx)
    S5 Data. Raw date of results in Fig 5.

    (XLSX)

    pbio.3003338.s011.xlsx (22.1KB, xlsx)
    S6 Data. Raw date of results in S1 Fig.

    (XLSX)

    pbio.3003338.s012.xlsx (14.8KB, xlsx)
    S7 Data. Raw date of results in S2 Fig.

    (XLSX)

    pbio.3003338.s013.xlsx (12KB, xlsx)
    S8 Data. Raw date of results in S4 Fig.

    (XLSX)

    pbio.3003338.s014.xlsx (13.2KB, xlsx)
    S9 Data. Raw date of results in S5 Fig.

    (XLSX)

    pbio.3003338.s015.xlsx (13.2KB, xlsx)
    S10 Data. Raw date of results in S6 Fig.

    (XLSX)

    pbio.3003338.s016.xlsx (16.1KB, xlsx)
    Attachment

    Submitted filename: Response to Reviewers.docx

    pbio.3003338.s019.docx (40.9KB, docx)
    Attachment

    Submitted filename: Response_to_Reviewers_auresp_3.docx

    pbio.3003338.s020.docx (40.9KB, docx)

    Data Availability Statement

    All relevant data are within the paper and its Supporting information files.


    Articles from PLOS Biology are provided here courtesy of PLOS

    RESOURCES