Introduction

The functional decline of stem cell activities in many tissues is a hallmark of aging. Intestinal stem cells (ISCs) are the only cell type capable of dividing in response to the physiological demands in the adult fly gut1,2. As in mammals, epithelial turnover in the Drosophila intestine is also maintained by ISCs. During homeostasis, ISCs are mostly quiescent but can initiate a proliferative response to stress conditions such as infection, toxins as well as aging3. During aging, the intestinal epithelium undergoes dysplasia, which is characterized by ISC over-proliferation and mis-differentiation, where ISCs show mixed feature with the enteroblast (EB), the daughter cell4,5.

Calcium is a versatile second messenger that is critical in regulating fundamental physiological functions. The concentration of Ca2+ varies between different subcellular compartments, which are established and maintained by Ca2+ channels and pumps6,7. The ER is the largest intracellular calcium store and Ca2+ homeostasis in ER is achieved by channels that release Ca2+ from the ER lumen into the cytosol - including IP3Rs and RyRs - and mechanisms that return Ca2+ to the ER, such as SERCA pumps and store-operated Ca2+ entry (SOCE)7,8. ER releases Ca2+ via IP3R and allows Ca2+ to passively enter the mitochondrial outer membrane (OMM) via VDAC1, a high conductance protein9. However, the inner mitochondrial membrane (IMM) is impermeable to Ca2+, and Ca2+ enters through the mitochondrial calcium uniporter (MCU) complex, which consists of the inner membrane channel MCU and regulatory subunits including MICU1 and EMRE10,11. High Ca2+ concentrations at the mitochondrial intermembrane space activate MCU to mediate the entry of Ca2+ into the mitochondrial matrix, where Ca2+ stimulates the activities of tricarboxylic acid (TCA) cycle, β-oxidation and respiratory complexes to enhance ATP generation12.

AMPK (AMPactivated protein kinase) is a cellular energy sensor that activated by falling energy status13,14. In mammalian cell lines, reduction of MitoCa2+ by disruption of MERCs or by the Ca2+ chelator BAPTA-AM induced a bioenergetic crisis and activates AMPK-mediated (macro)autophagy to cope with energy depletion15,16,17. Recently, we and others found that subcellular calcium level in cytosol (CytoCa2+) and mitochondria serves as critical coordinators during the transition of ISCs from quiescence to proliferation8,18,19. This adaption failed in old ISCs, which have sustained higher CytoCa2+ and lack efficient MitoCa2+ uptake and have undergone a Warburg-like metabolic switch involving increased lactate production and reduced ATP levels, contributing to their deregulated proliferative activity8,18. How MitoCa2+ was reduced and whether restoring MitoCa2+ can reverse the metabolic profile and age-related defects in ISCs remains largely unknown.

The MERCs, also known as mitochondria–associated membranes (MAMs), are the physical links between the ER and mitochondria, providing an interface for Ca2+ transfer9,20. These structures are dynamic and highly heterogeneous among tissues and species21. MERCs establishment is regulated by multiple scaffold and tethering factors that include among others MFN2, IP3Rs-GRP75-VDACs, FIS1-BAP31, VAPB-PTPIP51, Mmm1, and SIGMAR122,23. Increasing evidence indicated that MERC is one of the membrane sources for autophagosomes and is the initiation site for autophagy24. In multicellular organisms, the Atg1 complex (including ATG1, FIP200, Atg17 and Atg13) is targeted to the ER and recruits the class III (PI(3)P) kinase complex (including ATG6 and ATG14) for initiation and maturation of pre-autophagosomes upon autophagy induction25,26. For instance, the omegasome marker DFCP1/ZFYVE1 (double FYVE domain-containing protein 1) and ATG14 was recruited to the MERCs during starvation27. When MERCs were disrupted by MFN2 knockdown, ATG14 is no longer localized in the MERCs and the autophagosome formation is also inhibited28, suggesting that MERCs play an indispensable role for autophagosome formation. A recent study has also shown that IP3R mediated Ca2+ transients at the outer surface of the ER triggers autophagosome initiation by promoting the formation of fluid-like FIP200 puncta29. Whether the calcium oscillations generated between the ER and the cytosol, or between the ER and mitochondria via MERCs, are responsible for the initiation of autophagy, and how the oscillations are generated and maintained, remains unclear.

Accumulating evidence indicates that autophagy declines with age and that impaired autophagy is associated with stem cell aging30. In mice, the regenerative capacity of muscle stem cells is maintained by basal autophagy, and its deregulation leads to premature entry into an irreversible senescent state31. Similarly, autophagy genes such as Atg5 and Atg7 are necessary for ISC maintenance and gut homeostasis32. In Drosophila, autophagy deficient ISCs show elevated DNA damage and cell cycle arrest during aging and lead to unlimited ISC proliferation or apoptosis33,34. These results reveal the crucial role of autophagy in preserving proper stem cell function for the continuous renewal of the intestinal epithelium. Indeed, activation of AMPK is linked to improved intestinal homeostasis during aging and extended lifespan in Drosophila35. How autophagy was impaired during aging was largely unknown.

Here, we found that MitoCa2+ influx retrogradely activates IP3R in MERCs, generating Ca2+ oscillations in MERCs that promote autophagy independent of AMPK. Genetic restoration of mitoCa2+ levels re-establishes the MitoCa2+-IP3R-MERCs feedback loop and delays ISC aging by autophagy.

Results

Deficient MitoCa2+ uptake contributes to ISC dysfunction in aged Drosophila

Our previous results showed that MitoCa2+ was reduced in aged ISCs, while CytoCa2+ remained higher8,18. To further trace intracellular calcium pattern in ISCs during aging, genetically encoded calcium reporters targeted to mitochondria (UAS::MitoGCaMP3) or cytosol (UAS::GCaMP5) were expressed specifically in ISCs by combining esg::Gal4ts with Su(H)::Gal80 (ISC-Gal4ts). In ex vivo live-imaged midguts, CytoCa2+ level exhibited a progressive increase from around 30 days of age, whereas MitoCa2+ level decreased gradually (Fig. 1a and Supplementary Fig. 1a, b), implied that the breakdown of Ca2+ homeostasis started at middle age (The lifespan of flies under conventional conditions is estimated to be 60–70 days36). The simultaneous monitoring of CytoCa2+ and MitoCa2+ in individual ISCs (UAS::MitoGCaMP3 for mitochondria, UAS::RCaMP1 for cytosol) revealed that both CytoCa2+ and MitoCa2+ in ISCs from young animals (3 d old) were significantly increased upon infection using Erwinia carotovora carotovora 15 (Ecc15), a mild enteropathogen that induces a transient proliferative response in the gut epithelium37. In contrast, in ISCs from aged animals (30 d old), CytoCa2+ levels were further increased by infection, but MitoCa2+ levels remained lower than those in young animals (Fig. 1b). This suggested that the MitoCa2+ response to external stimuli is impaired in aged ISCs.

Fig. 1: Deficient MitoCa2+ uptake contributes to ISC dysfunction in aged Drosophila.
figure 1

a Plot summary of calcium level in cytosol and mitochondria at different time points. A, amplitude. b Representative images and quantification of MitoCa2+ (green) and CytoCa2+ (red) in ISCs of young (3 d) and old animals (30 d) with indicated genotypes exposed or not to Ecc15 infection. Data show mean ± S.E.M (n = 95, 122, 99 and 173 cells, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. Scale bar, 5 µm. c Representative images and quantification of relative MitoCa2+ levels in ISCs of young (3 d) or old animals (30 d) with indicated genotypes. MitoGCaMP3 (green), MitoDsRed (red). Data show mean ± S.E.M (n = 109, 110, 131, 104, 93 and 104 cells, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. Scale bar, 5 µm. d Quantification of relative MitoCa2+ levels in aged ISCs (30 d) with indicated genotypes exposed or not to Ecc15 infection. Data show mean ± S.E.M (n = 133, 129, 108 and 126 cells, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. e Quantification of the number of mis-differentiation ISCs with indicated genotypes at different ages. Data show mean ± S.E.M (n = 10, 8, 8 and 6 guts, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. f Schematics for MARCM clone analysis. Main cell types (upper left), clonal formation (clone of GFP positive cells) and clonal maintenance (shedding GFP positive cells in the clone) (bottom) were shown as schematics. g Representative images and quantification of clone numbers in the posterior midgut with indicated genotypes at different ages. ISC clones (green) and nuclei (blue). Data show mean ± S.E.M (n = 7, 9, 10, 8, 7 and 7 guts, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. *P  <  0.05 and ****P  <  0.0001. Scale bar: 50 µm. Each experiment was repeated for 3 times (be, g). See Supplementary Table 1 for genotypes. Source data are provided as a Source Data file.

The components of mitochondrial calcium uniporter (MCU) complex are conserved in Drosophila, such as the pore-forming MCU (CG18769) and the negative gatekeeper MICU1 (CG4495)38,39. The ISC specific overexpression of MCU (MCUOE) or the knockdown of MICU1 (MICU1RNAi) was sufficient to increase MitoCa2+ in young ISCs and also to restore MitoCa2+ in old ones (Fig. 1c and Supplementary Movie 1-6). Conversely, the overexpression of MICU1 (MICU1OE) or the knockdown of MCU (MCURNAi) was sufficient to decrease MitoCa2+ in young ISCs (Supplementary Fig. 1c and Supplementary Movie 7, 8). Notably, MCU overexpression also resulted in a reduction in elevated CytoCa2+ in aged ISCs (Supplementary Fig. 1a). Furthermore, MCU overexpression was sufficient to restore MitoCa2+ levels, which were reduced and insensitive to Ecc15 infection in aged ISCs (Fig. 1d and Supplementary Fig. 1d). These results suggested that MitoCa2+ levels in ISCs are controlled by the MCU complex.

We then examined whether genetically restoring MitoCa2+ was sufficient to rescue aged ISCs dysfunction. NRE-eGFP is a genetic reporter of Notch signaling40, which expressed exclusively in enteroblast cells (EB), while the percentage of NRE-eGFP positive ISCs significantly increased in aged midguts, indicating that the aged ISCs underwent mis-differentiation and displayed mixed feature of ISC and EB cells4, which can be restored by MCUOE (Fig. 1e and Supplementary Fig. 1e). Mosaic analysis with a repressible cell marker (MARCM) is a positive mosaic labeling system that has been widely used to study the self-renewal and differentiation abilities of ISCs in Drosophila1,2,41. Two different strategies were exploited to monitor ISC activity: clones were induced in young animals (3 d) and the clone number was checked 7days (3 + 7) or 25 days (3 + 25) after clone induction (ACI), reflecting the ability of ISCs pool to form and maintain these clones. Alternatively, clones were induced in aged animals (25 d) and then checked 7 days ACI to check the clone forming ability of aged ISCs (Fig. 1f). Intriguingly, both the clone formation and maintenance capacity of ISC declined in aged animals, and which can be restored by MCUOE (Fig. 1g). Taken together, these results indicated that old ISCs lack efficient MitoCa2+ uptake and the dysfunctions in aged ISCs can be improved by restoring MitoCa2+ levels.

Restoration of MitoCa2+ delays ISC aging and promotes intestinal homeostasis in aged flies

We then examined the role of MitoCa2+ in ISC aging and intestinal function. Cell cycle indicated by the FUCCI system (fluorescent ubiquitination-based cell cycle indicator)42 showed that the ratio of ISCs in G1 vs in G2/M stage was significantly increased in aged ISCs, indicating cell cycle dysregulation (Fig. 2a and Supplementary Fig. 2a). Consistent with previous studies18,34, lactate level (indicated by Laconic, a FRET based Lactate reporter) and DNA damages (indicated by gamma-H2AX staining) were increased in aged ISCs (Fig. 2b and Supplementary Fig. 2b, c). Intriguingly, cell cycle dysregulation, lactate level as well as gamma-H2AX accumulation were all restored in aged ISCs with MCUOE or MICU1RNAi by the TARGET system43 (Fig. 2a–c and Supplementary Fig. 2b–d) and also by a drug (RU486) inducible ISC/EB specific driver 5961::GS44 (Supplementary Fig. 2e). Conversely, fewer number and reduced clone size (cell number within the clone) were observed in MICU1OE (Supplementary Fig. 3a). Moreover, the knockdown of MCU in young ISCs (10 d) exhibited similarities to aged ISCs, including cell cycle dysregulation and DNA damage accumulations (Supplementary Fig. 3b, c). Age-related ISC dysfunction impairs midgut digestive functions, such as the loss of gut acid-base homeostasis and the decline of food intake and excretion45,46. Intriguingly, ISC-specific overexpression of MCU was sufficient to restore the deterioration of gut pH homeostasis (Fig. 2d) and mitigated the decrease of food intake (Fig. 2e) and excretion (Fig. 2f) in aged guts. Together, these results indicated that restoring MitoCa2+ levels is sufficient to delay ISC aging and restore gut homeostasis in Drosophila.

Fig. 2: Restoration of MitoCa2+ delays ISC aging and promotes intestinal homeostasis in aged flies.
figure 2

a Representative images and quantification of cell cycle stages (shown by different color) of ISCs with indicated genotypes at different ages by the FUCCI system. ISCs (green and/or red) and nuclei (blue). Data show mean ± S.E.M (n = 858, 1504, 978 and 1100 cells, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. Scale bar, 10 µm. b Quantification of relative lactate levels (Em 480/550 fluorescence ratios) in ISCs with indicated genotypes at different ages by Laconic, a FRET-based genetic reporter. Data show mean ± S.E.M (n = 234, 230, 256 and 226 cells, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. c The percentage of ISCs with the indicated number of γH2AvD foci in aged flies (30 d) with the indicated genotypes was quantified. Data show mean ± S.E.M (n = 340 and 229 cells, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. *P  <  0.05 and ****P  <  0.0001. d Gut homeostasis of flies with different genotypes at different ages were monitored by the pH indicator Bromophenol Blue. “Homeostasis”, a well-defined acidic Copper cell region is flanked by basic anterior midgut and posterior midgut; “Perturbed A,” the acidic region is lost and the whole gut is basic; “Perturbed B,” the strongly acidic region is lost, while the rest of the gut also becomes less basic (weak blue). Representative images shown on the bottom. Data show mean ± S.E.M (n = 15 guts per group) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. Scale bar: 1 mm. e Food intake measured using CAFÉ assay with indicated genotypes at different ages. Data show mean ± S.E.M (n = 12 flies per group) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. f Representative images and quantification of deposit with indicated genotypes at different ages. Data show mean ± S.E.M (n = 16 fields for each group of 15 flies) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. Scale bar, 1 mm. Each experiment was repeated for 3 times (af). See Supplementary Table 1 for genotypes. Source data are provided as a Source Data file.

Autophagy is activated by MitoCa2+ to delay ISC aging

MitoCa2+ level is essential for bioenergetics and our previous results showed that ATP level was reduced in aged ISCs18. Severe ATP depletion could lead to cell apoptosis or necrosis47. Previous studies have shown that genetic inhibition of β-oxidation leads to ISC necrosis48. Neutral lipid content (stained by LipidTOX) was largely unchanged by genetically inhibiting MitoCa2+ level in quiescent ISCs as well as in active ISCs upon infection (Supplementary Fig. 4a). The percentage of apoptotic (indicated by Cleaved Caspase-3 staining) or necrotic ISCs (indicated by Propidium iodide (PI) staining) was also largely unchanged in MCUOE or MICU1OE in aged ISCs (Supplementary Fig. 4b, c). We then tested whether MitoCa2+ delay ISC aging by regulating autophagy. Indeed, the number of autophagosomes (mCherry-Atg8 positive puncta) and acidic lysosomes (Lysotracker positive staining) in ISCs were significantly increased by MCUOE or by MICU1RNAi (Fig. 3a, b and Supplementary Fig. 5a, b). The mCherry-Atg8 positive vesicles in MCUOE ISCs readily colocalized with the lysosome (indicated by GFP-Lamp1), implied that these autophagosomes were readily fused with lysosome (Supplementary Fig. 5c). Lysosomal activity can be assessed using two fluorescent dye indicators: Magic Red Cathepsin B (Magic Red) and DQ-BSA49,50. More puncta of Magic Red and DQ-BSA were observed in ISCs by MCUOE (Fig. 3c, d and Supplementary Fig. 5d, e).

Fig. 3: Autophagy is activated by MitoCa2+ to delay ISC aging.
figure 3

ai Quantification of the number of mCherryATG8 (a, g), lysotracker (b, h), Magic Red (c), DQ-BSA (d), autophagosomes (f), autolysosomes (e, f) and Ref(2)P-GFP (i) positive puncta in young ISCs (3 d) with indicated genotypes and conditions, respectively. Data show mean ± S.E.M (n = 102, 111, 91 cells in a, n = 71, 66, 63 cells in b, n = 130, 96 cells in c, n = 76, 68 cells in d, n = 100, 100, 102 cells in e, n = 100, 101, 100, 94 cells in f, n = 112, 123, 147, 163, 140, 124 cells in g, n = 101, 122, 120, 105, 122, 106 cells in h and n = 75, 70, 54 cells in i, respectively, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test (a, b, ei) and two-tailed unpaired t-test (c, d). ****P  <  0.0001. j Representative images and quantification of the number of autophagosomes and/or autolysosomes in young ISCs (3 d) with indicated genotypes by TEM. VM, visceral muscles; M, mitochondria; N, nucleus. ISCs and their nucleus were circled in light yellow and typical autophagosomes or autolysosomes were highlighted by arrowheads. Data show mean ± S.E.M (n = 20, 21 and 20 cells, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. *P  <  0.05, **P  <  0.01 and ****P  <  0.0001. Scale bar, 1 µm. k Quantification of clone numbers in the posterior midgut with indicated genotypes at different ages after treated or not with CQ for indicated time. Data show mean ± S.E.M (n = 13, 11, 10, 12, 13, 11, 11 and 14 guts, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. l, m Quantification of the intensity of anti-γH2AvD signals in ISCs of aged flies (30 d) with indicated genotypes, respectively. Data show mean ± S.E.M (n = 8, 8, 9, 9 guts in l and n = 13, 9 guts in (m) from left to right) based on one-way ANOVA with Tukey’s multiple comparison test (l) and two-tailed unpaired t-test (m). ****P  <  0.0001. Each experiment was repeated for 3 times (am). See Supplementary Table 1 for genotypes. Source data are provided as a Source Data file.

Further track the autophagic flux by the tandem GFP-mCherry-Atg8a reporter in which GFP is rapidly quenched in an acidic environment51, showed that the percentage of autolysosomes (mCherry+/GFP-, red puncta) was significantly increased in MCUOE ISCs (Fig. 3e and Supplementary Fig. 5f). Consistently, accumulation of autophagosomes (mCherry+/GFP+, yellow puncta) were also increased in MCUOE ISCs under Chloroquine (CQ, an inhibitor of autophagosome-lysosome fusion52) treatment (Fig. 3f and Supplementary Fig. 5g). On the other hand, both of the increased mCherry-ATG8a- and Lysotracker-positive puncta induced by starvation were reduced in MCURNAi or MICU1OE ISCs (Fig. 3g, h and Supplementary Fig. 5h, i). Furthermore, although these autophagosomes can readily colocalize with GFP-Lamp1 (Supplementary Fig. 5c), GFP signals of GFP-mCherry-Atg8a were retained in MICU1OE ISCs (Supplementary Fig. 5f). Ref(2)P is the ortholog of P62 in Drosophila, which is a conserved autophagy substrate53,54. The number of Ref(2)P-GFP puncta in ISCs was significantly reduced by MCUOE, whereas it was increased by MICU1OE (Fig. 3i and Supplementary Fig. 5j). Transmission electron microscope (TEM) images also indicated that the number of autophagosomes and/or autolysosomes in ISCs was significantly increased by MCUOE or MICU1RNAi (Fig. 3j and Supplementary Fig. 5k). These data indicated that autophagic activity was sensitive to MitoCa2+ in ISCs of fly intestine. In addition, overexpression of MCU in indirect flight muscles by IFM::Gal4 also promoted autophagy (indicated by increased mCherryAtg8+ puncta) in muscles (Supplementary Fig. 6a). Treatment of CQ can completely block MCUOE induced ISC rejuvenation on the number of clones induced in old age (25 + 7) or maintenance of clones induced at young age (3 + 25) (Fig. 3k and Supplementary Fig. 6b). DNA damage suppressed by MICU1RNAi in aged ISCs was blocked when autophagy was disrupted by Atg1RNAi (Fig. 3l and Supplementary Fig. 6c). On the other hand, increasing autophagy by ISC specific Atg17 overexpression, a component in the ATG1 complex essential for phagophore formation55, was sufficient to suppress DNA damages in aged ISCs (Fig. 3m and Supplementary Fig. 6d-f). Together, these results indicated that MitoCa2+-mediated anti-aging effect is dependent on autophagy.

MitoCa2+ promotes autophagy by activating IP3R-mediated Ca2+ release from the ER

We then attempted to elucidate the mechanism by which MitoCa2+ stimulates autophagy. Autophagosome initiation in ISCs, as indicated by antibody staining against endogenous Atg8a56,57 (Fig. 4a), as well as by genetic reporters for Atg13 in the Atg1 complex, Atg6 in the PI(3)P kinase complex, Atg9 for autophagosome membrane source58 and ZFYVE1 in omegasomes, was robustly increased by MCUOE (Fig. 4b and Supplementary Fig. 7a-d). Endocytosis (indicated by the levels of Rab5 and Rab7, markers for early and late endosomes respectively), was mostly unaffected by MCUOE in ISCs (Supplementary Fig. 7e–g). AMPK is a central regulator of autophagy13. However, AMPK activity, indicated by AMPK-SPARK (a phase separation dependent genetic reporter of AMPK activity)59 and downstream mTOR activity (shown by phosphor-4EBP staining) was largely normal in MCUOE ISCs (Supplementary Fig. 8a, b). Consistently, the cell size of ISCs was largely unchanged by genetically regulating MitoCa2+ levels (Supplementary Fig. 8c). Meanwhile, although autophagic activity in ISCs (indicated by immune-staining against endogenous p6260) was found to be both sensitive to AMPK and MitoCa2+ levels, epistatic analysis indicated that the inhibition of AMPK activities suppresses autophagy mediated by MitoCa2+, and vice versa (Supplementary Fig. 8d). Excessive ROS level can activate autophagy through JNK(c-Jun N-Terminal Kinase)61. However, no obvious increase of mitochondrial ROS (H2O2) level (shown by a genetic ROS sensor MitoRoGFP2_Orp118), while JNK activity (indicated by TRE-RFP62) was even reduced by MCUOE in ISCs (Supplementary Figs. 8e, f).

Fig. 4: MitoCa2+ promotes autophagy by activating IP3R-mediated Ca2+ release from the ER.
figure 4

a Representative images and quantification of the number of autophagosomes in young ISCs (3 d, green) with indicated genotypes by anti-ATG8a staining (red). Data show mean ± S.E.M (n = 99 and 126 cells, from left to right) based on two-tailed unpaired t-test. ****P  <  0.0001. Scale bars, 5 μm. b Quantification of the number of ZFYVE1, ATG6, ATG9 and ATG13 positive puncta in young ISCs (3 d) with indicated genotypes. Data show mean ± S.E.M (n = 89, 112, 109, 65, 114, 89, 132 and 128 cells, from left to right) based on two-tailed unpaired t-test. ****P  <  0.0001. c Quantification of relative MitoCa2+ levels in young ISCs (3 d) with indicated genotypes. Data show mean ± S.E.M (n = 102, 123 and 118 cells, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. d Working scheme: MCU overexpression might activate IP3R (question mark) and cause Ca2+ release from ER, forming Ca2+ oscillations required for isolation membrane (IM) nucleation and autophagy initiation. e Representative images and quantification of relative ERCa2+ levels in young ISCs (3 d, red) with indicated genotypes by ER-GCaMP6 (green). Typical ISC (red) was circled by dashed lines. Data show mean ± S.E.M (n = 140, 115, 113 and 127 cells, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. Scale bars,10 μm. fh Quantification of the number of acidic lysosomes in young ISCs (3 d) with indicated genotypes, respectively. Data show mean ± S.E.M (n = 95, 82, 90, 94 cells in f, n = 95, 71, 94, 65 cells in g and n = 99, 92 cells in h, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test (f, g) and two-tailed unpaired t-test (h). ****P  <  0.0001. i, j Quantification of the intensity of anti-γH2AvD signals in aged ISCs (30 d) with indicated genotypes, respectively. Data show mean ± S.E.M (n = 13, 9, 10, 11 guts in i and n = 11, 9 guts in j, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test (i) and two-tailed unpaired t-test (j). ****P  <  0.0001. Each experiment was repeated for 3 times (ac, ej). See Supplementary Table 1 for genotypes. Source data are provided as a Source Data file.

IP3R, instead of RyR, is the main calcium release channel in the ER of Drosophila ISCs7,8 and Itpr encodes the sole homolog of IP3R in fly genome63. MitoCa2+ in ISCs was inhibited by Itpr knockdown (ItprRNAi), whereas increased by Itpr overexpression (ItprOE) (Fig. 4c and Supplementary Fig. 9a), indicating that MitoCa2+ level was coupled with IP3R-mediated Ca2+ release from the ER.

Since IP3R was shown to trigger Ca2+ oscillations on the ER outer membrane and induce autophagosome biogenesis29, we then explore whether MitoCa2+ can promote autophagy by eliciting Ca2+ oscillations on the ER outer membrane (Fig. 4d). ER-GCaMP6-210, a sensitive reporter of Ca2+ in the ER lumen (ERCa2+)64, was utilized to reflect the activity of IP3R. Surprisingly, increasing MitoCa2+ in ISCs by MICU1RNAi was sufficient to decrease ERCa2+ levels, which was blocked when Itpr was simultaneously silenced, indicating that increasing MitoCa2+ promoted Ca2+ release from the ER matrix by activating IP3R (Fig. 4e). Meanwhile, we generated an Itpr-HA knock-in transgene by CRISPR-Cas9 mediated homologous recombination (Supplementary Fig. 10a) to test whether its protein level or subcellular localization were altered by MitoCa2+ level. Indeed, Itpr-positive signals (indicated by anti-HA staining) form obvious puncta which are closely associated with KDEL-GFP (Supplementary Fig. 10b) in ISCs. Intriguingly, these puncta were significantly increased by MCUOE in ISCs, further indicated that MitoCa2+ promotes IP3R activity (Supplementary Fig. 10c). Autophagy induced by starvation or MICU1RNAi in ISCs (quantified by the number of autophagosomes and Lysotracker) was also diminished by ItprRNAi (Fig. 4f, g and Supplementary Fig. 9b–g), while overexpressing Itpr promoted ISCs autophagic activity (Fig. 4h and Supplementary Fig. 11a, b). On the other hand, DNA damage in aged ISCs suppressed by MICU1RNAi was blocked by ItprRNAi (Fig. 4i and Supplementary Fig. 11c), and ItprOE was sufficient to suppress DNA damage in aged ISCs (Fig. 4j and Supplementary Fig. 11d). These results suggested that an increase in mitochondrial Ca2+ promotes autophagy and delays ISCs aging by retrogradely activating Ca2+ release from the ER via IP3R.

MERCs-mediated calcium crosstalk between the ER and mitochondria involved in MitoCa2+-mediated autophagy in ISCs

MERCs as the physical links between the ER and mitochondria, not only serve as a ground for dynamic Ca2+ transfer from the ER to mitochondria65, but also for autophagic regulation28. We then tested whether MERCs is involved in the regulation of MitoCa2+-mediated autophagy. The physical contacts between mitochondria and ER, visualized by DsRed-targeted mitochondrial outer membrane (OMM) and KDEL-GFP labeled ER, were observed in ISCs and were enhanced by MCU overexpression (Supplementary Fig. 12a). To monitor MERCs in ISCs, we generated a transgenic fly expression MAMtracker-Green (MAM-GFP), a dimerization-dependent GFP (ddGFP) based reporter66, to detect reversible changes in MERCs structure under the control of UAS-Gal4 system. When overexpressed by ISC-Gal4ts, the MAM-GFP positive puncta were adjacent and partially colocalized with mitochondria and ER and autophagosomes (Supplementary Fig. 12b–e). Meanwhile, the number of MAM-GFP positive vesicles in ISCs is sensitive to genetic manipulations of known MERCs components. For instance, genetically knocking down Grp75 (CG8542), or Mmm1 (CG13838, a conserved component of MAM also participate in phospholipid trafficking in yeast67) was sufficient to reduce MERCs structure, while overexpressing Grp75 or human VAPB (hVAPB) was sufficient to increase MERCs (Fig. 5a), indicating that the MERCs structures are controlled by the conserved tethering complex in Drosophila and MAM-GFP is a sensitive tool to quantify the MERCs dynamics in ISCs. As expected, MitoCa2+ in ISCs was reduced when MERCs were genetically disrupted by Grp75RNAi or Mmm1RNAi, but increased when MERCs were augmented by Grp75 or hVAPB overexpression (Fig. 5b and Supplementary Fig. 13a), indicating that MitoCa2+ levels are controlled by MERCs in ISCs. Surprisingly, increasing MitoCa2+ by MCUOE promoted MERCs formation, as indicated by the number of MAM-GFP and structures observed under TEM microscope (Fig. 5c–e and Supplementary Fig. 13b). Moreover, Itpr knockdown profoundly suppressed the number of MAM-GFP puncta induced by MICU1RNAi, whereas ItprOE was sufficient to increase MERCs (Fig. 5f and Supplementary Fig. 13c), indicating that increasing MitoCa2+ promoted MERCs structures by activating IP3R. Intriguingly, autophagy induced by MitoCa2+ was suppressed by Mmm1RNAi or Grp75RNAi (Fig. 5g, h and Supplementary Fig. 14a, b), while increasing MERCs by overexpressing Grp75 or hVAPB was sufficient to promote autophagy (Fig. 5i, j and Supplementary Fig. 14c, d). Increasing MERCs by expressing an artificial tether, Tether-RFP68, was also sufficient to increase autophagy in ISCs as indicated by Atg8a positive puncta (Fig. 5k). Taken together, these data indicated that MERCs-mediated calcium crosstalk is critical for MitoCa2+-mediated autophagy in ISCs.

Fig. 5: MERCs-mediated calcium crosstalk between the ER and mitochondria involved in MitoCa2+-mediated autophagy in ISCs.
figure 5

a Representative images and quantifications of the number and intensity of MAM-GFP positive vesicles (green) in young ISCs (3 d) with indicated genotypes. Typical ISCs (red) were circled in dashed lines. Data show mean ± S.E.M (n = 98, 107, 104, 121 and 107 cells, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. Scale bars,5 μm. b Quantification of relative MitoCa2+ levels in young ISCs (3 d) with indicated genotypes. Data show mean ± S.E.M (n = 123, 136, 104, 108 and 111 cells, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. c Quantification of the number of MAM-GFP positive vesicles in young ISCs (3 d) with indicated genotypes. Data show mean ± S.E.M (n = 129 and 127 cells, from left to right) based on two-tailed unpaired t-test. ****P  <  0.0001. d, e Representative TEM images (d) and quantifications (e) of MERCs in young ISCs (3 d) with indicated genotypes. Typical ISCs nucleus (yellow). Data show mean ± S.E.M (n = 8 and 8 sections, from left to right) based on two-tailed unpaired t-test (e). **P  <  0.01. Scale bars,2 μm (d). f Quantification of the number of MAM-GFP positive vesicles in young ISCs (3 d) with indicated genotypes. Data show mean ± S.E.M (n = 124, 114, 97, 100 and 95 cells, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. gj Quantification of the number of mCherryATG8a (g, i) and lysotracker (h, j) positive puncta in young ISCs (3 d) with indicated genotypes. Data show mean ± S.E.M (n = 69, 75, 85, 97 cells in g, n = 93, 81, 88, 81 cells in h, n = 131, 115, 102 cells in (i) and n = 90, 79, 75 cells in (j), from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. k Representative images and quantification of the number of autophagosomes in young ISCs (3 d) with indicated genotypes by anti-ATG8a staining (magenta). ISCs (green). Data show mean ± S.E.M (n = 76 and 66 cells, from left to right) based on two-tailed unpaired t-test. ****P  <  0.0001. Scale bars, 5 μm. Each experiment was repeated for 3 times (ak). See Supplementary Table 1 for genotypes. Source data are provided as a Source Data file.

MERCs are critical for MitoCa2+-mediated rejuvenation of aged ISCs

We then investigated the role of MERCs in MitoCa2+-mediated rejuvenation of aged ISCs. Genetical upregulation of MERCs was sufficient to restore MitoCa2+ levels, which were reduced and insensitive to Ecc15 infection in aged ISCs (Fig. 6a, b and Supplementary Fig. 15a, b). Furthermore, DNA damages in aged ISCs suppressed by MCUOE were blocked by Mmm1RNAi (Fig. 6c and Supplementary Fig. 16a). Lactate levels suppressed by MICU1RNAi in aged ISCs were also increased by Grp75RNAi (Fig. 6e and Supplementary Fig. 16d). Intriguingly, DNA damages, lactate levels and mis-differentiation were all restored in aged ISCs by hVAPBOE or Grp75OE (Fig. 6d–f and Supplementary Fig. 16b–d). Moreover, overexpression of Tether-RFP was also sufficient to reduce DNA damages in aged ISCs (Fig. 6g). Taken together, these data indicated that MERCs are critical for MitoCa2+-mediated rejuvenation of aged ISCs.

Fig. 6: MERCs are critical for MitoCa2+-mediated rejuvenation of aged ISCs.
figure 6

a, b Quantification of relative MitoCa2+ levels in aged ISCs (30 d) with indicated genotypes exposed or not to Ecc15 infection. Data show mean ± S.E.M (n = 100, 120, 95 cells in (a) and n = 137, 128, 110, 92 cells in (b), from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. c, d Quantification of the intensity of γH2AvD signals in aged ISCs (30 d) with indicated genotypes. Data show mean ± S.E.M (n = 11, 9, 9, 12 guts in (c) and n = 10, 8 guts in d, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test (c) and two-tailed unpaired t-test (d). ****P  <  0.0001. e Quantification of relative lactate levels in aged ISCs (30 d) with indicated genotypes by Laconic, a FRET-based genetic reporter. Data show mean ± S.E.M (n = 104, 94, 93, 102 and 102 cells, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. f Quantification of the number of mis-differentiation ISCs in aged flies (30 d) with indicated genotypes. Data show mean ± S.E.M (n = 7 and 11 guts, from left to right) based on two-tailed unpaired t-test. ****P  <  0.0001. g Representative images and quantification of the intensity of anti-γH2AvD signals (magenta) in aged ISCs (30 d) with indicated genotypes. ISCs (green), nuclei (blue). Data show mean ± S.E.M (n = 10 and 14 guts, from left to right) based on two-tailed unpaired t-test. ****P  <  0.0001. Scale bars, 10 μm. Each experiment was repeated for 3 times (ag). See Supplementary Table 1 for genotypes. Source data are provided as a Source Data file.

Breakdown of ER-mitochondria calcium crosstalk triggers ISCs aging due to defective autophagy

We then explore how MitoCa2+ was reduced in aged ISCs. Expression levels of components in the MCU complex, such as MCU, MICU1 and EMRE, were largely unchanged in the whole intestine at different ages, as shown by RT-qPCR (Supplementary Fig. 17a). Furthermore, we found that the expression levels of some calcium-associated channels (PMCA, Stim, Orai, IP3R, RyR, VDAC, MCU, MICU1 and EMRE) and mitochondria-ER contact tethers (Mfn2, Grp75, VAPB, Mmm1) were largely unchanged in old ISCs compared to young ISCs by RNA-seq dataset analysis (GSE157794, Supplementary Fig. 17b). Then we hypothesized that dysfunctional IP3R and/or MERCs lead to decreased MitoCa2+ in aged ISCs. Intriguingly, compared to young ISCs, despite no significant changes in ERCa2+ levels, ERCa2+ responded poorly to starvation in aged ISCs (Fig. 7a and Supplementary Fig. 17c), indicating that the activity of IP3R was impaired in aged ISCs. Meanwhile, the number and intensity of MAM-GFP puncta in aged ISCs were lower than those in young ISCs under both feeding and starvation conditions (Fig. 7b, c and Supplementary Fig. 17d), suggesting that the structure of MERCs was damaged in aged ISCs. These results showed that the breakdown of IP3R-mediated Ca2+ crosstalk in MERCs is critical for MitoCa2+ reduction in aged ISCs.

Fig. 7: The breakdown of the ER-mitochondria calcium crosstalk triggers ISCs aging due to defective autophagy.
figure 7

a Quantification of relative ERCa2+ levels in ISCs with indicated genotypes and conditions at different ages. NF, normal food; Stv, starvation. Data show mean ± S.E.M (n = 150, 161, 109, 122, 159 and 110 cells, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. b, c Quantifications of the number and intensity of MAM-GFP positive vesicles in ISCs with indicated genotypes and conditions at different ages. Data show mean ± S.E.M (n = 116, 120, 121, 124, 128 and 117 cells, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. *P  <  0.05 and ****P  <  0.0001. d, e Quantifications of the number of Ref(2)P-GFP (d) and mCherryATG8 (e) positive puncta in ISCs with indicated conditions and genotypes at different ages. Data show mean ± S.E.M (n = 84, 61, 78, 79 cells in (d) and n = 123, 93, 78, 154, 147, 138 cells in (e), from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. f Quantification of the intensity of AMPKSPARK signals in ISCs with indicated genotypes at different ages. Data show mean ± S.E.M (n = 96, 93, 107 and 110 cells, from left to right) based on two-way ANOVA with Tukey’s multiple comparison test. ***P  < 0.001 and ****P  <  0.0001. g Quantification of the number of mCherryATG8a positive puncta in aged ISCs (30 d) with indicated genotypes. Data show mean ± S.E.M (n = 96, 74 and 70 cells, from left to right) based on one-way ANOVA with Tukey’s multiple comparison test. ****P  <  0.0001. h Model schematic: in young animal ISCs, calcium oscillations in MERC were maintained by the positive cycle formed by MitoCa2+-IP3R-MERC. In short, increasing MitoCa2+ influx by MCU overexpressing lead to the activation of IP3R, which then promote calcium release from the ER and formed oscillations in MERC which is maintained by complexes such as VAPB-PTPIP51 and IP3R-VDAC for the initiation of autophagy independent of AMPK. In aged ISCs, the MitoCa2+/IP3R/MERC cycle was disrupted, leading to impaired autophagy. Autophagosomes are indicated by double membrane vesicles. Each experiment was repeated for 3 times (ag). See Supplementary Table 1 for genotypes. Source data are provided as a Source Data file.

On the other hand, more p62 aggregates were observed in aged ISCs than those in young ones (Fig. 7d and Supplementary Fig. 18a, b). The number of mCherryATG8a+ puncta in aged ISCs were lower than those in young ISCs under both feeding and starvation conditions (Fig. 7e and Supplementary Fig. 18c). Notably, unlike young ISCs, starvation failed to induce an increase of mCherryATG8a+ puncta in aged ISCs (Fig. 7e and Supplementary Fig. 18c). These results showed that autophagy was impaired and autophagic machinery was less sensitive to starvation in aged ISCs. Surprisingly, although autophagic activity was reduced, AMPK activity (indicated by AMPK-SPARK positive puncta) was actually increased in aged ISCs (Fig. 7f and Supplementary Fig. 19a), which may be due to ATP level drop caused by reduced MitoCa2+ in aged ISCs18. Indeed, increasing MitoCa2+ by overexpressing MCU downregulated AMPK activity in aged ISCs (Fig. 7f and Supplementary Fig. 19a), suggesting that autophagic failure in aged ISCs is not due to the suppression of AMPK.

Interestingly, increasing MitoCa2+ by overexpressing MCU was sufficient to restore IP3R activity, MERCs as well as autophagy in aged ISCs (Fig. 7a–e, and Supplementary Fig. 17c, d, 18a–c). Similarly, overexpression of Itpr or hVAPB was sufficient to promote autophagic activity in aged ISCs (Fig. 7g and Supplementary Fig. 19b). These results indicated that MitoCa2+, IP3R and MERCs form a positive feedback loop and the breakdown of the loop triggers ISC aging through autophagic failure, even with elevated AMPK activity (Fig. 7h).

Discussion

Inter-organelle communication in eukaryotic cells is spatially and temporally organized and display a transient dynamic pattern responds to pathophysiological conditions69. MERCs, as the platforms between ER and mitochondria, actively involved in processes such as lipid trafficking, calcium crosstalk, redox homeostasis, and autophagy70. This is corroborated by the growing body of evidence indicating that MERCs play a pivotal role in regulating the aging process and age-related diseases71. However, the mechanisms governing the number, composition, and function of MERCs, as well as the factors that disrupt their regulation during aging, remain largely unelucidated.

Here we found that subcellular Ca2+ homeostasis is disturbed in ISCs of aged flies: Ca2+ levels in the ER lumen and cytosol are maintained at higher levels, while those in the mitochondria are reduced. Compromised autophagy is a common feature of aging across diverse species72. In mammalian cell lines, the pro- or anti-autophagic effect of Ca2+ depends on the context, kinetics, amplitude or spatial extent. For example, many studies have shown that increased Cytosolic Ca2+ signals can activate AMPK through CaMKKβ by phosphorylating pThr17273,74. Activated AMPK then promotes autophagy by inhibiting mTORC1 or controversially by phosphorylating ULK175,76. While other studies have also shown that the elevation of CytoCa2+ increases protein phosphatase4, which inactivates AMPK through dephosphorylation of pThr17277. Reduced MitoCa2+ can activate AMPK due to energy deprivation78. A recent study showed that a temporal decrease in ATP during mitosis would activate AMPK, which can activate MCU through phosphorylation79. In this scenario, the AMPK/MCU couple helps to maintain MitoCa2+ level and energy status. However, we found AMPK activity was even increased in aged ISCs, the impaired autophagy and poor sensitivity to starvation was due to the disruption of the positive feedback loop consisting of MitoCa2+, MERCs via IP3R. Our results further indicated that AMPK and the MitoCa2+/IP3R/MERCs cascade are mutually required for autophagy induction. These results indicated that these two modules function in parallel to regulate autophagy induction in ISCs.

Based on the results, we propose a model in which the Ca2+ homeostasis in MERCs is maintained by a balance between Ca2+ release from the ER and Ca2+ uptake by mitochondria. Forced MitoCa2+ uptake would cause temporal drop of Ca2+ in MERCs, which activates IP3R and cause replenishment of Ca2+ in MERCs and forms calcium oscillations (Fig. 7h). How IP3R detect local Ca2+ drops in MERCs? IP3Rs are unusually large proteins, comprising tetramers of closely-related subunits, and each with about 2700 amino acid residues80. In addition to IP3, Ca2+ was shown also to regulate IP3R with a biphasic concentration dependence81; modest increases in CytoCa2+ levels enhancing responses to IP3, while higher levels of CytoCa2+ were inhibitory for IP3R82. It’s possible that forced MitoCa2+ influx cause Ca2+ drop in the microdomain, which was sensed by IP3R and lead to its activation. IP3R activity is also regulated by its subcellular localization and post-translational modifications83. For instance, Cdk5, a kinase localizes in MERCs, phosphorylates IP3R1 at Ser421 and regulates ER Ca2+ transfer to the mitochondria in mouse embryonic fibroblasts (MEFs)84. Therefore, MitoCa2+ may also regulate IP3R activity through local Ca2+-dependent kinases or phosphatases in MERCs. Here, we found that Ca2+ oscillations in MERCs mediated by mitochondrial Ca2+ and IP3R form a positive feedback loop to regulate autophagy. Breakdown of this loop is a driven force for ISC aging and genetically restoring the loop is sufficient led to ISC rejuvenation through AMPK-independent autophagy. Given that the components are conserved, it is plausible that the MERCs-specific Ca2+ oscillations also function as a module for autophagy regulation in mammals. Our results provide a strategy to tackle aging and aging-related diseases.

Methods

Fly strains and reagents

Flies were cultured at 25 °C and 60% humidity with a 12 h light/dark cycle on standard yeast/molasses-based food unless otherwise indicated (recipe for 1 L food: 13.8 g agar, 22 g molasses, 75 g malt extract, 18 g dry inactivated yeast, 80 g corn flour, 10 g soy flour, 6.25 ml, propionic acid, 2 g methyl 4-hydroxybenzoate, 7.2 mL ethanol and water up to 1 L).

The following strains were obtained from the Bloomington Drosophila Stock Center: w1118 (BL3605), tub::GAL80ts (BL7108), UAS-mCherry (BL25774), UAS-nls-GFP (BL4775), UAS-tdTomato (BL36327), UAS-MitoDsRed (BL93056), UAS-20XUAS-RCaMP1f (BL52210), UAS-tdTomato-P2A-GCaMP5G (BL80079), UAS-MCU (BL16357), FRT40A (BL5759), UAS-GFP.E2f1.1-230,UAS-mRFP1.CycB.1-266 (BL55099), NRE::EGFP.S (BL30727), UAS-MitoRoGFP2_Orp1 (BL67667), UAS-GFP-mCherry-Atg8a (BL37749), UAS-mCherry-ATG8a (BL37750), UAS-ItprRNAi (BL25937), UAS-Itpr (BL30742), UAS-ER.GCaMP6-210 (BL83293), UAS-Hsc70-5.FLAG.HA.V (BL95230), UAS-hVAPB.HA (BL82230), UAS-GFP-Lamp1 (BL42714), UAS-GFP-Rab5 (BL43336), UAS-GFP-Rab7 (BL42705), UAS-ATG17(BL91226), UAS-GFP-KDEL (BL9899), UAS-RFP-KDEL (BL30909), UAS-AMPKCA (BL32110), UAS-AMPKDN (BL32112). UAS-MICU1RNAi (v110456), UAS-MCURNAi (v110781), UAS-Hsc70-5RNAi (v47745) were obtained from Vienna Drosophila Stock Center (VDRC). UAS-MICU1.ORF.3xHA (F000962) were obtained from FlyORF. UAS-ATG1RNAi (THU2357), UAS-mCD8-GFP (THJ0080), UAS-Mmm1RNAi (THU5442) were obtained from TsingHua University Fly Center. esg::GAL4 was a gift from S.Hayashi, Gbe::GAL80 from S.X.Hou, UAS-MitoGCaMP3 from F.Kawasaki, hsFlp;FRT40,tub::GAL80;tub::GAL4,UAS-GFP from B.Ohlstein, UAS-Laconic from O.Morris, UAS-GFP-Ref(2)P from T.Neufeld, TRE-RFP from D.Bohmann, UAS-AMPK-SPARK from H.Huang, UAS-GFP-ATG6, UAS-GFP-ATG9, UAS-GFP-ZFYVE1, UAS-Tether-RFP from C.Tong, UAS-mCherry-ATG13 from S.J.He. A full list of fly genotypes used in each figure panel is available in Supplementary Table 1.

Drosophila TARGET system, aging and MARCM induction

We used the TARGET system43 and the Gene Switch system44 for conditional expression of UAS-linked transgenes. Conditional expression in ISCs + EBs was achieved by combining ISC/ EB driver esg::Gal4 (esg::G4) with Gal80ts (G80ts), a temperature-sensitive Gal4 inhibitor driven by a tubulin promoter (Tub::G80ts). For conditional expression in only ISCs, esg::Gal4 and Tub::Gal80ts were combined with Su(H)GBE-G80, which drives the expression of a temperature-insensitive version of Gal80 in EBs. When using the TARGET system, flies were cultured at permissive temperature (18 °C) to avoid transgene expression before being shifted to the restrictive temperature of 29 °C for the indicated time to induce transgene expression. For experiments using our aging model, flies were kept at 18 °C for the indicated time before shifting to 29 °C for 4 days to induce transgene expression. Food was changed every 2–3 days.

For experiments with the 5961 GeneSwitch system, flies were raised for 3 days on normal food post eclosion and then kept for 4 days on food supplemented with RU486 at 5 mg/ml (mifepristone; Calbiochem #475838) dissolved in ethanol, or with ethanol alone (control).

For MARCM clone induction, flies were heat-shocked at 37 °C for 1 h. Following heat-shock, flies were kept at 25 °C and dissected after the indicated time. For starvation treatments, flies were transferred to fresh vials containing 5% sucrose at 29 °C for 8 h prior to dissection. In all experiments, only female posterior midgut was analyzed.

Cloning and transgenic fly generation

MAMtracker-Green (MAM-GFP) was amplified from pMAMtracker-Green, which was a gift from Koji Yamanaka, using primers Fw: GAATTCatggctattcaactcagg and Rev: CTCGAGttacttcatgaaaagacc. Amplified fragments were then cloned into pUAST-attB vector using EcoRI and XhoI restriction enzyme sites. For Itpr-HA knock-in transgenic fly, the CRISPR-Cas9-based strategy was utilized to generate the homology-mediated knock-in allele of Itpr (CG1063, the fruit fly IP3R homolog). Briefly, the sgRNA sequence in the CDS of CG1063 was designed by CRISPOR and cloned into the pEASY-Blunt vector. Three set of primers were designed for the donor plasmid: the primer pair for the homologous arms spanning the stop codon of CG1063 and the primer for the linker plus 3xHA sequence. The sequences were then sequentially cloned into the pMD™18-T vector. In this way, the donor plasmid was knocked into the CG1063 coding sequence just before the stop codon in frame. The sgRNA and donor construct were then microinjected into fly (BL78781) embryos. Positive transgenic lines were verified by sequencing.

The primers for sgRNA CG1063 sg-F 5’ tacagactcactgaaacggcGTTTTAGAGCTAGAAATAGC 3’ and CG1063 sg-R 5’ gccgtttcagtgagtctgtaCGACGTTAAATTGAAAATAGG 3’. The primers for donors CG1063 Lpart-FCGACGATATCTCTAGctaattaaaattataactgactttcagttactttaccact, CG1063 Lpart-RAGTCCGGActgaaacggcaggaggc, CG1063 linker+3x HA-F gtttcagTCCGGACTCAGATCTGGCAGC, CG1063 linker+3x HA-R cagactcaAGCGTAGTCTGGGACGTCG, CG1063 Rpart-F CTACGCTtgagtctgtaagaaatagaatagaatactaattttggcac and CG1063 Rpart-R ATGACCATGATTACGCCTTCAATCCACTCCACGC. The transformation injection was performed in the Core Facility of Drosophila Resource and Technology, CAS (Shanghai).

Ecc15 infection and Chloroquine treatment

Bacteria Ecc15 were cultured overnight at 30 °C in LB medium. A concentrated bacterial pellet (A600 nm of ~200), which was centrifuged from 1-ml overnight culture media, was resuspended 1:1 with 5% sucrose to yield an Ecc15 infection solution. Flies were starved at 29 °C for 2 h and then transferred to fresh vials covered with a filter paper (WhatmanTM) and either 150 μL of Ecc15 infection solution or 5% sucrose (mock) solution. Flies were infected for 16 h before dissection for Ca2+ imaging.

As previously described85, for CQ treatment (Chloroquine diphosphate salt solid, #C6628, Sigma-Aldrich), tubes containing food supplemented with 100 μM CQ were used. Briefly, for MARCM clones, flies were starved at 25 °C for 2 h to synchronize feeding and transferred to tubes containing food supplemented with 100 μM CQ for 2 days before heat-shock. Following heat-shock, flies were kept at 25 °C and transferred to tubes containing supplemented 100 μM CQ every 2days and dissected after the indicated time.

Bromophenol blue assay

The bromophenol blue assay was used to detect the pH change in the midgut as previously described45. Briefly, 100 µl of 2% Bromophenol blue sodium (Sigma, B5525) was added to the food surface and several holes were poked to allow the Bromophenol blue solution completely absorbed by the food. Images were taken immediately after 16 h.

Cafe assay and fly excretion measurement

For Café assay, a 5 µl capillary containing liquid food (10% yeast, 10% sucrose, and blue dye) was inserted into a cotton plug of 6 cm long vial, and 1 cm high 1% agar was at the vial bottom to keep the moisture. The food intake was recorded every 12 h and the capillary was changed every 24 h. For fly excretion measurement, flies were dry starved for 2 h and put into Bromophenol blue vial food for 24 h, and then the deposits on the vial wall were imaged and quantified.

RNA isolation and quantitative real-time PCR

RNA was isolated from guts of 15 flies using TriZol reagent (#15596026 Life technologies). Around 20 ug RNA then were reversely transcribed using 5X All-In-One RT Mastermix With Accurt Kit (#G592, Applied Biological Materials) according to the manufacturer’s instructions. All of the reactions were performed in triplicate in a CFX96TM Real-Time System (Bio-Rad Laboratories). The following primers pairs were used: MCU, 5′-GTCTCGCCCTGCGTTTGG-3′ and 5′-CGAAGCTTCTGTGCTGCTG-3′; MICU1, 5′-GTGGCCATGGTCAATCTTTC-3′ and 5′-TTGTTGCTGAGTTGGTTGTCA-3′; EMRE, 5′-ACATGTCCAGCGTGTACTTTC-3′and 5′-GGTATGACGGCACAGAAGATG-3′; Rp49, 5′-GCCGCTTCAAGGGACAGTATCTG-3′and 5′-AAACGCGGTTCTGCATGAG-3′. Expression was normalized with Rp49 and quantification was performed using the comparative CT method.

Transmission electron microscopy

Guts were dissected in PBS, fixed with 2.5% glutaraldehyde, and washed three times with 0.1 M phosphate buffer at 4 °C, post-fixed with 1.0% OsO4 for 1 h at 4 °C, washed with phosphate buffer, dehydrated in ethanol, and embedded in Spurr resin. Thick section was stained with Toluidine blue and ultra-thin sections of the embedded guts were double-stained with uranyl acetate and lead citrate and examined with a JEM1400 transmission electron microscope (TEM).

Immunostaining and florescence microscopy

Female fly midguts were dissected in PBS and fixed for 20 min at room temperature in 4% formaldehyde. After fixation, the samples were washed with PBS containing 0.5% bovine serum albumin and 0.1% Triton X-100 (PBST) for 1 h. Appropriate primary antibodies were added to the samples and incubated at 4 °C overnight. Midguts were washed and then incubated with secondary antibodies for 4 h at room temperature. Samples were mounted in Mounting Medium (#S2100, Solarbio). The following antibodies or dyes were used: rabbit anti-Histone H2AvD pS137 (1:200; #600-401-914, Rockland); rabbit anti-Ref(2)P (1:500; #ab178440, Abcam); rabbit anti-Atg8a (GABARAP57) (1:500, #13733S Cell Signaling Technology); rabbit anti-HA (1:250; #3724 T, Cell Signaling); rabbit anti-cleaved Caspase 3 (1:100; #Y407558, Applied Biological Materials); rabbit anti-phospho-4EBP1 (1:200; #2855S, Cell Signaling); rabbit anti-GFP (1:500; #PABG1, Chromotek); Alexa-488-conjugated donkey anti-rabbit secondary antibody (1:1000;#A21206, Molecular Probes); Goat-α-rabbit-IgG(H  +  L)-Cy3 secondary antibody (1:1000; #A10520, Molecular Probes); Goat-α-rabbit-IgG(H  +  L)-Cy5 secondary antibody (1:1000; #A10523, Molecular Probes); and Hoechst 33342 (0.1 μg ml-1, 6 min incubation; #H1399, Invitrogen). All images were taken on a Zeiss Axio Imager M2 fluorescence microscope equipped with a 40× Plan-Apochromat oil-immersion objective lens (Zeiss) or an Olympus IXplore SpinSR confocal microscope equipped with a 100×1.5 NA oil-immersion objective lens (Olympus) and processed using Adobe Photoshop, Illustrator and Image J.

LipidTOX, Lysotracker, propidium iodide, Magic Red and DQ-BSA staining

LipidTox staining was performed as previously described86. In brief, fixed guts were washed 3 times for 5 min in PBST, then directly stained with LipidTOX™ solution (1:200, # H34476, Invitrogen) diluted in PBST, mounted in Mounting Medium with Hoechst and analyzed on a confocal microscope.

For Lysotracker staining, guts were dissected in AHLS and incubated in AHLS containing 0.5 μM Lysotracker Red (#C1046, Beyotime) for 5 min at room temperature. They were then mounted similarly to all other live-imaging experiments and imaged immediately by a two-photon laser microscope.

For propidium iodide staining, guts were dissected in PBS and then stained in 1.5 μM PI (#P1304MP, Invitrogen) for 15 min at room temperature. The guts were then fixed for 20 min in 4% formaldehyde, washed three times in PBS, mounted in Mounting Medium with Hoechst and analyzed on a confocal microscope.

For Magic Red and DQ-BSA staining, guts were dissected in AHLS and incubated in AHLS containing Magic Red (1:250, #ICT-941, Immunochemistry) or DQ-BSA (4 μg ml–1, #D12051, Invitrogen) for 10 min at room temperature. They were then mounted similarly to all other live-imaging experiments and imaged immediately by a two-photon laser microscope.

Ex vivo Drosophila intestine imaging setup

Flies were dissected in Adult Hemolymph-like Saline (AHLS) culture media containing 2 mM CaCl2, 5 mM KCl, 5 mM HEPES, 8.2 mM MgCl2, 108 mM NaCl, 4 mM NaHCO3, 1 mM NaH2PO4, 5 mM Trehalose and 10 mM Sucrose. Guts were rapidly transferred to a microscope slide with thin adhesive spacers (#GBL654008, Sigma-Aldrich), embedded in 1% low melting agarose (in AHLS) and submerged in AHLS. All live-imaging experiments were imaged using an Olympus FVMPE-RS two-photon microscope equipped with two lasers for fluorescence excitation. Only the middle posterior midgut region (region R4) was analyzed.

Image acquisition and quantifications

For Cyto[Ca2+], Mito[Ca2+] and ER[Ca2+] imaging experiments except Fig. 1B, identical sized z stacks covering a single layer of the Drosophila epithelium were imaged using an Olympus FVMPE-RS two-photon microscope equipped with two lasers for fluorescence excitation. An InSight X3-OL laser (Spectra-Physics) tuned to 530–550 nm was used to excite TdTomato/DsRed/mCherry, and a MaiTai eHPDS-OL laser (Spectra-Physics) tuned to 488 nm was used to excite GCaMP3/5/6. The laser beams were focused using a 25× 1.05 NA water-immersion objective lens (Olympus). GCaMP3/5/6 emission was captured at 495–540 nm and TdTomato/DsRed/mCherry emission at 575–645 nm. For Cyto[Ca2+] and Mito[Ca2+] imaging experiments in Fig. 1b, images were taken on a Zeiss Axio Imager M2 fluorescence microscope equipped with an apotome. The laser beams were focused using a 40× Plan-Apochromat oil-immersion objective lens (Zeiss).

For the ratiometric MitoRoGFP2_Orp1 biosensor, experiments were performed based previous publications18,87. Freshly dissected midgut tissues were incubated with 20 mM NEM (N-ethyl maleimide, #E387, Sigma-Aldrich) (in AHLS) for 10 min before imaging. Probe fluorescence was excited sequentially using the 405 nm and 488 nm laser lines and emission was detected at 495–540 nm.

For Laconic FRET imaging experiments, following excitation using the 440 nm laser line, emission was sequentially detected at 460–500 nm and 520–560 nm.

For AMPK-SPARK imaging experiments, EGFP was excited at 488 nm and emission was detected at 495–540 nm, mCherry was excited at 545 nm and emission was detected at 575–645 nm.

All quantifications were done manually using FIJI (NIH ImageJ) software. When sensors were expressed using esg::G4 in the absence of Su(H)GBE::G80ts, regions of interest (ROIs) were manually drawn so as to only segment smaller cells that likely represent ISCs and early-stage EBs. Larger esg::G4 positive cells, which are probably late-stage EBs that are soon approaching differentiation into ECs, were excluded from analysis.

For MARCM clone analysis, the clone formation and maintenance capacity of ISCs was measured by quantifying the number of MARCM clones containing at least 3 cells in the posterior midgut.

For ISC cell cycle analysis using the FUCCI system, the nucleus of GFP, RFP, GFP/RFP and non-fluorescent labeled ISC indicated cell residing in G1, S, G2 and G0 phase, respectively. The number of different fluorochrome-tagged nuclei was manually quantified across the whole posterior midgut.

For ISC differentiation analysis, ISCs and EB cells were labeled with mCherry and GFP, respectively. The mis-differentiation of ISCs was measured by quantified the percentage of ISCs labeled with mCherry and GFP.

For γH2AvD measurements, γH2AvD foci per nuclei of ISCs was counted manually. Mean intensity quantifications were made in individual ISCs of GFP channels normalized to nearby background.

For p-4EBP1, Ref(2)P, TRE-RFP and GFP-MAM measurements, mean intensity quantifications were made in individual ISCs of GFP or mCherry channels normalized to nearby background.

For mCherry-ATG8a, Lysotracker, GFP-Ref(2)P, GFP-ZFYVE1, GFP-ATG6, GFP-ATG9, mCherry-ATG13 and GFP-MAM puncta measurements, number quantifications were counted in individual ISCs of GFP or mCherry channels.

For Cyto[Ca2+], Mito[Ca2+] and ER[Ca2+] measurements, mean intensity quantifications were made in the ROIs of RCAMP1, MitoGCaMP3, GCaMP5/TdTomato, MitoGCaMP3/TdTomato, MitoGCaMP3/MitoDsRed or ERGCaMP6/mCherry channels.

For ratiometric measurements of MitoRoGFP2_Orp1, mean intensity quantifications were made in the ROIs of both 405 nm and 488 nm excitation channels.

For ratiometric FRET measurements of Laconic, mean intensity quantifications were then made in the ROIs of both the 460–500 nm and 520–560 nm emission channels.

For ratiometric measurements of AMPKSPARK, mean intensity quantifications were made in the ROIs of EGFP/mCherry channels.

For Magic Red and DQ-BSA staining, the number and intensity of puncta were quantified in ISCs which labeled by GFP.

Statistical analysis

Statistical analysis is performed using GraphPad Prism 8. Data derived from different genetic backgrounds or different conditions were compared using Student’s two-tailed unpaired t-test, one-way ANOVA followed by Tukey’s multiple comparison test, two-way ANOVA followed by Tukey’s multiple comparison test as indicated in the figure legends. All data were presented as mean ± SEM. The significance was set at P  <  0.05. *P  <  0.05, **P  <  0.01, ***P  <  0.001, ****P  <  0.0001, ns = non-significant (P  >  0.05). All experiments were performed in biological triplicates.

Reporting summary

Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.