Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Adaptor protein AP-3 produces synaptic vesicles that release at high frequency by recruiting phospholipid flippase ATP8A1

Abstract

Neural systems encode information in the frequency of action potentials, which is then decoded by synaptic transmission. However, the rapid, synchronous release of neurotransmitters depletes synaptic vesicles (SVs), limiting release at high firing rates. How then do synapses convey information about frequency? Here, we show in mouse hippocampal neurons and slices that the adaptor protein AP-3 makes a subset of SVs that respond specifically to high-frequency stimulation. Neurotransmitter transporters slot onto these SVs in different proportions, contributing to the distinct properties of release observed at different excitatory synapses. Proteomics reveals that AP-3 targets the phospholipid flippase ATP8A1 to SVs; loss of ATP8A1 recapitulates the defect in SV mobilization at high frequency observed with loss of AP-3. The mechanism involves recruitment of synapsin by the cytoplasmically oriented phosphatidylserine translocated by ATP8A1. Thus, ATP8A1 enables the subset of SVs made by AP-3 to release at high frequency.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: AP-3 targets VAMP7 and a subset of VGLUT1 to low release probability SVs.
Fig. 2: AP-3 maintains the EPSCs evoked by high-frequency stimulation.
Fig. 3: Proteomic comparison of WT and mocha SVs identifies phospholipid flippase ATP8A1.
Fig. 4: Loss of AP-3 redistributes ATP8A1 away from SVs.
Fig. 5: Loss of ATP8A1 impairs release at high frequency.
Fig. 6: Loss of ATP8A1 reduces the SV association of synapsin.
Fig. 7: Interaction with synapsin mediates the function of ATP8A1 in transmitter release.
Fig. 8: ATP8A1 maintains the EPSCs evoked by high-frequency stimulation.

Similar content being viewed by others

Data availability

The original mass spectra for the SV proteomics are shown in Supplementary Table 1. The sequence of primers are shown in Supplementary Table 2. Source data are provided with this paper.

Code availability

The code used for the analysis of the pHluorin imaging is provided with this paper as Supplementary Code 1.

References

  1. Schultz, W. Getting formal with dopamine and reward. Neuron 36, 241–263 (2002).

    Article  CAS  PubMed  Google Scholar 

  2. O’Keefe, J. Place units in the hippocampus of the freely moving rat. Exp. Neurol. 51, 78–109 (1976).

    Article  PubMed  Google Scholar 

  3. Zucker, R. S. & Regehr, W. G. Short-term synaptic plasticity. Ann. Rev. Physiol. 64, 355–405 (2002).

    Article  CAS  Google Scholar 

  4. Südhof, T. C. The presynaptic active zone. Neuron 75, 11–25 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Jackman, S. L., Turecek, J., Belinsky, J. E. & Regehr, W. G. The calcium sensor synaptotagmin 7 is required for synaptic facilitation. Nature 529, 88–91 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Alabi, A. A. & Tsien, R. W. Synaptic vesicle pools and dynamics. Cold Spring Harb. Perspect. Biol. 4, a013680 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Kavalali, E. T. The mechanisms and functions of spontaneous neurotransmitter release. Nat. Rev. Neurosci. 16, 5–16 (2015).

    Article  CAS  PubMed  Google Scholar 

  8. Pereira, D. B. et al. Fluorescent false neurotransmitter reveals functionally silent dopamine vesicle clusters in the striatum. Nat. Neurosci. 19, 578–586 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Silm, K. et al. Synaptic vesicle recycling pathway determines neurotransmitter content and release properties. Neuron 102, 786–800 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Takei, K., Mundigl, O., Daniell, L. & De Camilli, P. The synaptic vesicle cycle: a single vesicle budding step involving clathrin and dynamin. J. Cell Biol. 133, 1237–1250 (1996).

    Article  CAS  PubMed  Google Scholar 

  11. Watanabe, S. et al. Clathrin regenerates synaptic vesicles from endosomes. Nature 515, 228–233 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Kononenko, N. L. et al. Clathrin/AP-2 mediate synaptic vesicle reformation from endosome-like vacuoles but are not essential for membrane retrieval at central synapses. Neuron 82, 981–988 (2014).

    Article  CAS  PubMed  Google Scholar 

  13. Faúndez, V., Horng, J. T. & Kelly, R. B. A function for the AP3 coat complex in synaptic vesicle formation from endosomes. Cell 93, 423–432 (1998).

    Article  PubMed  Google Scholar 

  14. Li, P., Merrill, S. A., Jorgensen, E. M. & Shen, K. Two clathrin adaptor protein complexes instruct axon-dendrite polarity. Neuron 90, 564–580 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Nakatsu, F. et al. Defective function of GABA-containing synaptic vesicles in mice lacking the AP-3B clathrin adaptor. J. Cell Biol. 167, 293–302 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Salazar, G. et al. The zinc transporter ZnT3 interacts with AP-3 and it is preferentially targeted to a distinct synaptic vesicle subpopulation. Mol. Biol. Cell 15, 575–587 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Scheuber, A. et al. Loss of AP-3 function affects spontaneous and evoked release at hippocampal mossy fiber synapses. Proc. Natl Acad. Sci. USA 103, 16562–16567 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Vogt, K., Mellor, J., Tong, G. & Nicoll, R. The actions of synaptically released zinc at hippocampal mossy fiber synapses. Neuron 26, 187–196 (2000).

    Article  CAS  PubMed  Google Scholar 

  19. Voglmaier, S. M. et al. Distinct endocytic pathways control the rate and extent of synaptic vesicle protein recycling. Neuron 51, 71–84 (2006).

    Article  CAS  PubMed  Google Scholar 

  20. Evstratova, A., Chamberland, S., Faundez, V. & Toth, K. Vesicles derived via AP-3-dependent recycling contribute to asynchronous release and influence information transfer. Nat. Commun. 5, 5530 (2014).

    Article  CAS  PubMed  Google Scholar 

  21. Miesenböck, G., De Angelis, D. A. & Rothman, J. E. Visualizing secretion and synaptic transmission with pH-sensitive green fluorescent proteins. Nature 394, 192–195 (1998).

    Article  PubMed  Google Scholar 

  22. Hua, Z. et al. v-SNARE composition distinguishes synaptic vesicle pools. Neuron 71, 474–487 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Mori, Y., Takenaka, K.-I., Fukazawa, Y. & Takamori, S. The endosomal Q-SNARE, syntaxin 7, defines a rapidly replenishing synaptic vesicle recycling pool in hippocampal neurons. Commun. Biol. 4, 981 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Ramirez, D. M., Khvotchev, M., Trauterman, B. & Kavalali, E. T. Vti1a identifies a vesicle pool that preferentially recycles at rest and maintains spontaneous neurotransmission. Neuron 73, 121–134 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Ashrafi, G., Wu, Z., Farrell, R. J. & Ryan, T. A. GLUT4 mobilization supports energetic demands of active synapses. Neuron 93, 606–615 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Galli, T. et al. A novel tetanus neurotoxin-insensitive vesicle-associated membrane protein in SNARE complexes of the apical plasma membrane of epithelial cells. Mol. Biol. Cell 9, 1437–1448 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Fremeau, R. T. Jr. et al. The expression of vesicular glutamate transporters defines two classes of excitatory synapse. Neuron 31, 247–260 (2001).

    Article  CAS  PubMed  Google Scholar 

  28. Fremeau, R. T. Jr. et al. Vesicular glutamate transporters 1 and 2 target to functionally distinct synaptic release sites. Science 304, 1815–1819 (2004).

    Article  CAS  PubMed  Google Scholar 

  29. Manabe, T. & Nicoll, R. A. Long-term potentiation: evidence against an increase in transmitter release probability in the CA1 region of the hippocampus. Science 265, 1888–1892 (1994).

    Article  CAS  PubMed  Google Scholar 

  30. Kantheti, P. et al. Mutation in AP-3 δ in the mocha mouse links endosomal transport to storage deficiency in platelets, microsomes and synaptic vesicles. Neuron 21, 111–122 (1998).

    Article  CAS  PubMed  Google Scholar 

  31. Salazar, G., Craige, B., Love, R., Kalman, D. & Faundez, V. Vglut1 and ZnT3 co-targeting mechanisms regulate vesicular zinc stores in PC12 cells. J. Cell Sci. 118, 1911–1921 (2005).

    Article  CAS  PubMed  Google Scholar 

  32. Newell-Litwa, K., Salazar, G., Smith, Y. & Faundez, V. Roles of BLOC-1 and adaptor protein-3 complexes in cargo sorting to synaptic vesicles. Mol. Biol. Cell 20, 1441–1453 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Taoufiq, Z. et al. Hidden proteome of synaptic vesicles in the mammalian brain. Proc. Natl Acad. Sci. USA 117, 33586–33596 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. van der Velden, L. M. et al. Heteromeric interactions required for abundance and subcellular localization of human CDC50 proteins and class 1 P4-ATPases. J. Biol. Chem. 285, 40088–40096 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Coleman, J. A., Vestergaard, A. L., Molday, R. S., Vilsen, B. & Andersen, J. P. Critical role of a transmembrane lysine in aminophospholipid transport by mammalian photoreceptor P4-ATPase ATP8A2. Proc. Natl Acad. Sci. USA 109, 1449–1454 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Lee, S. et al. Transport through recycling endosomes requires EHD1 recruitment by a phosphatidylserine translocase. EMBO J. 34, 669–688 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Paterson, J. K. et al. Lipid specific activation of the murine P4-ATPase Atp8a1 (ATPase II). Biochemistry 45, 5367–5376 (2006).

    Article  CAS  PubMed  Google Scholar 

  38. Levano, K. et al. Atp8a1 deficiency is associated with phosphatidylserine externalization in hippocampus and delayed hippocampus-dependent learning. J. Neurochem. 120, 302–313 (2012).

    Article  CAS  PubMed  Google Scholar 

  39. Kay, J. G., Koivusalo, M., Ma, X., Wohland, T. & Grinstein, S. Phosphatidylserine dynamics in cellular membranes. Mol. Biol. Cell 23, 2198–2212 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Benfenati, F., Greengard, P., Brunner, J. & Bähler, M. Electrostatic and hydrophobic interactions of synapsin I and synapsin I fragments with phospholipid bilayers. J. Cell Biol. 108, 1851–1862 (1989).

    Article  CAS  PubMed  Google Scholar 

  41. Murray, J., Cuccia, L., Ianoul, A., Cheetham, J. J. & Johnston, L. J. Imaging the selective binding of synapsin to anionic membrane domains. Chembiochem 5, 1489–1494 (2004).

    Article  CAS  PubMed  Google Scholar 

  42. Benfenati, F. et al. Synaptic vesicle-associated Ca2+/calmodulin-dependent protein kinase II is a binding protein for synapsin I. Nature 359, 417–420 (1992).

    Article  CAS  PubMed  Google Scholar 

  43. Orenbuch, A. et al. Synapsin selectively controls the mobility of resting pool vesicles at hippocampal terminals. J. Neurosci. 32, 3969–3980 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Milovanovic, D., Wu, Y., Bian, X. & De Camilli, P. A liquid phase of synapsin and lipid vesicles. Science 361, 604–607 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Chi, P., Greengard, P. & Ryan, T. A. Synapsin dispersion and reclustering during synaptic activity. Nat. Neurosci. 4, 1187–1193 (2001).

    Article  CAS  PubMed  Google Scholar 

  46. Rosahl, T. W. et al. Essential functions of synapsins I and II in synaptic vesicle regulation. Nature 375, 488–493 (1995).

    Article  CAS  PubMed  Google Scholar 

  47. Stefani, G. et al. Kinetic analysis of the phosphorylation-dependent interactions of synapsin I with rat brain synaptic vesicles. J. Physiol. 504, 501–515 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Chi, P., Greengard, P. & Ryan, T. A. Synaptic vesicle mobilization is regulated by distinct synapsin I phosphorylation pathways at different frequencies. Neuron 38, 69–78 (2003).

    Article  CAS  PubMed  Google Scholar 

  49. Verstegen, A. M. J. et al. Phosphorylation of synapsin I by cyclin-dependent kinase-5 sets the ratio between the resting and recycling pools of synaptic vesicles at hippocampal synapses. J. Neurosci. 34, 7266–7280 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. De Gois, S. et al. Identification of endophilins 1 and 3 as selective binding partners for VGLUT1 and their co-localization in neocortical glutamatergic synapses: implications for vesicular glutamate transporter trafficking and excitatory vesicle formation. Cell. Mol. Neurobiol. 26, 679–693 (2006).

    Article  CAS  PubMed  Google Scholar 

  51. Farsad, K. et al. Generation of high curvature membranes mediated by direct endophilin bilayer interactions. J. Cell Biol. 155, 193–200 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Schuske, K. R. et al. Endophilin is required for synaptic vesicle endocytosis by localizing synaptojanin. Neuron 40, 749–762 (2003).

    Article  CAS  PubMed  Google Scholar 

  53. Vinatier, J. et al. Interaction between the vesicular glutamate transporter type 1 and endophilin A1, a protein essential for endocytosis. J. Neurochem. 97, 1111–1125 (2006).

    Article  CAS  PubMed  Google Scholar 

  54. Weston, M. C., Nehring, R. B., Wojcik, S. M. & Rosenmund, C. Interplay between VGLUT isoforms and endophilin A1 regulates neurotransmitter release and short-term plasticity. Neuron 69, 1147–1159 (2011).

    Article  CAS  PubMed  Google Scholar 

  55. Park, D. et al. Cooperative function of synaptophysin and synapsin in the generation of synaptic vesicle-like clusters in non-neuronal cells. Nat. Commun. 12, 263 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Gitler, D. et al. Different presynaptic roles of synapsins at excitatory and inhibitory synapses. J. Neurosci. 24, 11368–11380 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Denker, A. et al. A small pool of vesicles maintains synaptic activity in vivo. Proc. Natl Acad. Sci. USA 108, 17177–17182 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Hosaka, M., Hammer, R. E. & Südhof, T. C. A phospho-switch controls the dynamic association of synapsins with synaptic vesicles. Neuron 24, 377–387 (1999).

    Article  CAS  PubMed  Google Scholar 

  59. Kook, S. et al. AP-3-dependent targeting of flippase ATP8A1 to lamellar bodies suppresses activation of YAP in alveolar epithelial type 2 cells. Proc. Natl Acad. Sci. USA 118, e2025208118 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Sato, M. et al. The role of VAMP7/TI-VAMP in cell polarity and lysosomal exocytosis in vivo. Traffic 12, 1383–1393 (2011).

    Article  CAS  PubMed  Google Scholar 

  61. Roghani, A., Shirzadi, A., Butcher, L. L. & Edwards, R. H. Distribution of the vesicular transporter for acetylcholine in the rat central nervous system. Neuroscience 82, 1195–1212 (1998).

    Article  CAS  PubMed  Google Scholar 

  62. Bellocchio, E. E. et al. The localization of the brain-specific inorganic phosphate transporter suggests a specific presynaptic role in glutamatergic transmission. J. Neurosci. 18, 8648–8659 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Goh, G. Y. et al. Presynaptic regulation of quantal size: K+/H+ exchange stimulates vesicular glutamate transport. Nat. Neurosci. 14, 1285–1292 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Guan, S., Price, J. C., Prusiner, S. B., Ghaemmaghami, S. & Burlingame, A. L. A data processing pipeline for mammalian proteome dynamics studies using stable isotope metabolic labeling. Mol. Cell Proteomics 10, M111.010728 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Clauser, K. R., Baker, P. & Burlingame, A. L. Role of accurate mass measurement (± 10 ppm) in protein identification strategies employing MS or MS/MS and database searching. Anal. Chem. 71, 2871–2882 (1999).

    Article  CAS  PubMed  Google Scholar 

  66. Conn, C. S. et al. The major cap-binding protein eIF4E regulates lipid homeostasis and diet-induced obesity. Nat. Metab. 3, 244–257 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Orlando, M. et al. Functional role of ATP binding to synapsin I in synaptic vesicle trafficking and release dynamics. J. Neurosci. 34, 14752–14768 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Xu, H. et al. SNX5 targets a monoamine transporter to the TGN for assembly into dense core vesicles by AP-3. J. Cell Biol. 221, e202106083 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank J. Maas and other members of the Edwards laboratory for helpful discussion, J. Maas, X. Chen and K. Bender for help with electrophysiology, and R. Nicoll for help with the experimental design and evaluating the data. All confocal images were acquired at the UCSF Center for Advanced Light Microscopy-Nikon Imaging Center supported by an NIH S10 Shared Instrumentation grant (no. 1S10OD017993-01A1). MS was provided by the MS Resource at UCSF (A.L. Burlingame, Director) supported by the Dr. Miriam and Sheldon G. Adelson Medical Research Foundation and the UCSF Program for Breakthrough Biomedical Research. This work was supported by NIH grant nos. R01 MH50712 and R01 NS103938 to R.H.E. The funders had no role in study design, data collection, analysis, decision to publish or preparation of the paper.

Author information

Authors and Affiliations

Authors

Contributions

H.X., S.J. and R.H.E. conceptualized the work. M.K. performed the initial SV preparations. J.A.O.-P. performed the proteomics analysis with support from A.B. J.L. prepared several of the primary neuronal cultures and ATP8A1-related constructs. H.X. and R.H.E. wrote the paper.

Corresponding author

Correspondence to Robert H. Edwards.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Neuroscience thanks the anonymous reviewers for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 VAMP7+ and VGLUT2+ SVs differ in the frequency dependence of exocytosis.

a, Hippocampal neurons from WT mice were immunostained for endogenous synaptophysin (green), VAMP7 (red) and MAP2 (blue). The area inside the white rectangle is shown below at higher magnification. Arrowheads indicate punctae double labeling for synaptophysin and VAMP7. Scale bars, 20 μm. Representative images were from three independent experiments. b, Hippocampal neurons expressing VAMP7-, VAMP2- or VGLUT2-pHluorin (pH) were stimulated at 5, 10, 25 and 50 Hz for 20 s and the fluorescence normalized to total pHluorin as shown in Fig. 1. Graphs show the initial rate of fluorescence increase per action potential (P = 2.3 × 0−14, 5 and 10 Hz; 3.7 × 10−14, 25 Hz; 1.44 × 10−4, 50 Hz). n = 20/3 c, Graph shows the time constant for endocytosis of VAMP7- and VGLUT2-pH (P = 1.15 × 10−13, 5 Hz; 9.3 × 10−14, 10 Hz; 2.19 × 10−10, 25 Hz; 1.85 × 10−10, 50 Hz). n = 18/3 d-f, Hippocampal neurons expressing VAMP7- or VGLUT2-pH were stimulated at 5, 10, 25 and 50 Hz for 60-120 s in the presence of bafilomycin (d). The initial rate of fluorescence increase per action potential (e) and the proportion of pHluorin reporters (f) as a function of stimulation frequency. e, P = 3.9 × 10−14, 5 Hz; 4.1 × 10−14, 10 Hz; 1.11 × 10−13, 25 Hz; 0.27, 50 Hz (n = 18/3). f, P = 3.4 × 10−14, 5 Hz; 3.41 × 10−14, 10 Hz; 3.4 × 10−14, 25 Hz; 0.007, 50 Hz (n = 23/3). g, The initial rate/AP at 50 Hz normalized to the rate at 5 Hz in the presence or absence of bafilomycin: without bafilomycin (n = 20/3), P = 8.66 × 10−6, VGLUT2; 1.75 × 10−7, VAMP7; with bafilomycin (n = 18/3), P = 9.86 × 10−6, VGLUT2; 4.86 × 10−8, VAMP7. n = x/y where x is the number of fields examined and y the number of independent experiments. The data indicate mean ±  s.e.m. for individual coverslips containing 50–100 boutons each. **, P < 0.01; ***, P < 0.001; ****, P < 0.0001 by two-way (b,c,e,f) or one-way ANOVA (g) with Tukey’s multiple comparisons test.

Source data

Extended Data Fig. 2 VAMP7+ SVs differ from canonical SVs in pH, Ca++ sensitivity and coupling to presynaptic Ca++ channels.

a, Hippocampal neurons expressing VGLUT1-, VGLUT2-, VAMP2- or VAMP7-pH were imaged in Tyrode’s solution and 2-methanesulfonic acid (MES, pH 5.5) was added to quench cell surface pHluorin, then 50 mM NH4Cl to alkalinize the intracellular pool and reveal the total pHluorin fluorescence. Left, imaging; middle, fluorescence traces; right, surface expression of pHluorin reporters (P = 1.0 × 10−15, n = 19/3). Scale bar, 20 µm. b, The baseline intracellular fluorescence determined by subtraction of the cell surface protein and normalization to total in NH4Cl was used to determine the lumenal pH for each SV reporter (P = 1.92 × 10−10, n = 10/3). c-f, Hippocampal neurons from WT mice expressing VAMP7- (c), VGLUT2- (e) or VGLUT1-pH (f) were stimulated at 5, 10, 25 and 50 Hz in 0, 0.5, 2 and 4 mM Ca++, and the fluorescence response normalized as above (n = 15/3 for c and e, 20/3 for f and 17/3 for d). Graph shows the time constant for endocytosis at 50 Hz (d) (P = 0.027). g Hippocampal neurons expressing VAMP7- or VGLUT2-pH were incubated with 100 μM EGTA-AM for 15 min in Tyrode’s buffer and then stimulated at 10 Hz for 20 s. Scatterplot shows the peak response in EGTA-AM relative to controls (P = 2 × 10−15, n = 23/3). h, Hippocampal neurons expressing GLUT4-, VAMP7- or VGLUT1-pH were incubated with 1 mM AICAR for 30 min and the fluorescence response normalized as above. The scatter plot indicates the pH of GLUT4+ vesicles, determined from the response to acid quenching and neutralization with NH4Cl (n = 33/3). n = x/y where x is the number of fields examined and y the number of independent experiments. The data indicate mean ±  s.e.m. for individual coverslips containing 50-100 boutons each. ****, P < 0.0001 by one-way ANOVA with Tukey’s multiple comparisons test (a,b,d) or two-tailed unpaired t test (g).

Source data

Extended Data Fig. 3 VAMP7-pH and VGLUT2-pH exocytosis differ in dependence on VAMP2.

a, Hippocampal neurons from WT mice expressing VAMP7-, VGLUT1-, VMAT2- or VGLUT2-pH were treated with tetanus toxin (TeNT) for 16-18 hours and then stimulated at 5, 10, 25 and 50 Hz, with the fluorescence normalized as in Fig. 1. Graphs indicate fluorescence at the end-stimulation. VAMP7, P = 1.73 × 10−9, 10 Hz; 3.56 × 10−10, 25 Hz; 3.56 × 10−10, 50 Hz (n = 12/3). VGLUT1, P = 1.008 × 10−9,10 Hz; 4.0 × 10−14, 25 Hz; 4.0 × 10−14, 50 Hz (n = 15/3). VMAT2, P = 0.0012, 10 Hz; 7.22 × 10−9, 25 Hz; 3.56 × 10−10, 50 Hz (n = 12/3). VGLUT2, P = 0.02, 5 Hz; 6.57 × 10−8, 10 Hz; 3.56 × 10−10, 25 Hz and 50 Hz (n = 12/3). b, Hippocampal neurons from WT or VAMP2 KO mice expressing VAMP7- or VGLUT2-pH were stimulated at 5, 10, 25 and 50 Hz and the fluorescence response normalized as above. Graphs indicate the response at the end-stimulation. VAMP7, P = 5.12 × 10−9, 10 Hz; 1.08 × 10−7, 25 Hz; 3.56 × 10−10, 50 Hz (n = 12/3). VGLUT2, P = 3.98 × 10−5, 5 Hz; 3.56 × 10−10, 10 Hz; 3.56 × 10−10, 25 Hz; 3.56 × 10−10, 50 Hz (n = 12/3). c, Hippocampal neurons from WT or VAMP7 KO mice expressing VGLUT1-, VMAT2- or VGLUT2-pH were treated with TeNT for 16-18 hours and then stimulated at 50 Hz, with the fluorescence normalized as above. Graphs indicate fluorescence at the end-stimulation. VGLUT1, P = 1.05 × 10−7 (n = 15/3). VMAT2, P = 6.40 × 10−13 (n = 13/3). VGLUT2, P = 0.39 (n = 15/3). d, Hippocampal neurons from embryonic synaptotagmin 1 (Syt1) KO and littermates expressing VAMP7- or VGLUT2-pH were stimulated at 5, 10, 25 and 50 Hz. Graphs indicate the peak response. VAMP7, P = 1.14 × 10−4, 25 Hz; 2.61 × 10−7, 50 Hz (n = 11/3). VGLUT2, P = 0.0015, 5 Hz; 1.13 × 10−6, 10 Hz; 3.79 × 10−12, 25 Hz; 8.44 × 10−7, 50 Hz (n = 11/3). n = x/y where x is the number of fields examined and y the number of independent experiments. The data indicate mean ±  s.e.m. *, P < 0.05; **, P < 0.01; ****, P < 0.0001 by two-way ANOVA (a,b,d) or one-way ANOVA (c) with Tukey’s multiple comparisons test.

Source data

Extended Data Fig. 4 Loss of AP-3 specifically impairs release at high frequency.

a, Hippocampal neurons from WT and mocha mice expressing VAMP7-pH were stimulated at 5, 10, 25 and 50 Hz for 60-150 s in the presence of 60 nM bafilomycin and the fluorescence response normalized as shown in Fig. 1 (above). Statistical analysis of the response at 4 s is shown below. 25 Hz: P = 6 × 10−15, 2 s; 1.22 × 10−12, 3 s; 9 × 10−14, 4 s. 50 Hz: P = 2.16 × 10−7, 1 s; 5.71 × 10−11, 2 s; 2.01 × 10−12, 3 s; 2.9 × 10−14, 4 s). n = 16/3. b, The initial rate of fluorescence increase per action potential in WT and mocha neurons: P = 0.029,10 Hz; 1.133 × 10−6, 25 Hz; 3.56 × 10−10, 50 Hz (n = 12/3). c, The proportion of VAMP7-pH that responds to stimulation (recycling pool size) in WT and mocha neurons: P = 0.0084, 10 Hz; 1.12 × 10−13, 25 Hz; 5.2 × 10−14, 50 Hz (n = 16/3). n = x/y where x is the number of fields examined and y the number of independent experiments. The data indicate mean ±  s.e.m. for individual coverslips containing 50-100 boutons each. *, P < 0.05; **, P < 0.01; ****, P < 0.0001 by two-way ANOVA with Tukey’s multiple comparisons test.

Source data

Extended Data Fig. 5 Effect of the mocha mutation on short-term plasticity in CA1 strata radiatum and lacunosum-moleculare.

a, Hippocampal region CA1 with whole cell recording electrode in the pyramidal cell layer. For VGLUT1+ inputs, stimulating electrodes were placed in stratum radiatum between regions CA3 and CA1. For VGLUT2+ inputs, stimulating electrodes were placed in stratum lacunosum-moleculare between regions CA3 and CA1. b, Representative traces from three independent experiments in response to stimulation at 1 and 25 Hz from stratum radiatum of WT and mocha slices. c, Representative traces from three independent experiments in response to stimulation at 5, 25 and 50 Hz from stratum lacunosum-moleculare of WT and mocha slices. d, Analysis of the normalized stratum lacunosum-moleculare EPSCs at 10, 30 and 90 action potentials as a function of stimulation frequency (n = 11 cells from three WT mice; n = 14 cell for 5 and 50 Hz, 9 cells for 25 Hz from three mocha mice). There is no significant difference between the genotypes by two-way ANOVA with Tukey’s multiple comparisons test. The data indicate mean ±  s.e.m.

Source data

Extended Data Fig. 6 Proteomic comparison of SVs from WT and mocha mice.

a, Diagram of SV purification. After velocity sedimentation in glycerol, fractions 5-9 (pink) were collected, sedimented and analyzed by LC-MS/MS. b, Enrichment of SV proteins. Equal amounts of protein (0.5 µg) from the different stages of SV purification (H, homogenate; P1, tissue debris, nuclei, and large myelin fragments; P2, synaptosomes; LP1, synaptic plasma membrane and associated organelles; LS2, synaptosomal cytoplasm; LP2, synaptic vesicles) were analyzed by immunoblotting. Representative blots were from three independent experiments. c, Fractionation of synaptic vesicles (LP2) by glycerol velocity sedimentation showing the migration of VAMP7, VGLUT2 and synaptophysin. Representative blots were from three independent experiments. d, Fold enrichment of proteins in mocha SVs relative to WT. e, Quantitative western analysis of SV proteins in WT and mocha mice. Scatterplots show the fraction relative to WT. P = 4.24 × 10−5, VAMP7; 0.04, Syt1, 0.0019, VGLUT1; 8.05 × 10−6, ZnT-3; 3.24 × 10−3, ATP8A1; 0.0052, Scamp1 (n = 3 independent experiments). The data indicate mean ±  s.e.m. *, P < 0.05; **, P < 0.01; ***, P < 0.001; ****, P < 0.0001 by unpaired two-tailed t-test.

Source data

Extended Data Fig. 7 Loss of AP-3 redistributes ATP8A1 away from axon terminals in hippocampus and cerebellum.

a, Slices from adult WT and mocha mice were immunostained for ATP8A1 (red), synaptophysin (green), Hoechst (blue) and MAP2 (purple). Scale bar, 400 μm. Representative images were from three independent experiments. b,c, Cerebellar slices from adult WT and mocha mice were immunostained for ATP8A1 (red), synaptophysin (green) and Hoechst (blue). Images were acquired at low (b) and high magnification (c). Scale bars, 400 μm (b) and 100 μm (c). Scatterplot indicates the fluorescence intensity of ATP8A1 from the molecular layer of WT and mocha slices: b, P = 3.36 × 10−6; c, P = 2.69 × 10−8, ATP8A1; c, 0.9809, synaptophysin (15 images from 3 independent experiments). The data indicate mean ±  s.e.m. ****, P < 0.0001 by unpaired two-tailed t-test.

Source data

Extended Data Fig. 8 Loss of ATP8A1 selectively impairs the response to high frequency stimulation.

a, Quantitative western analysis of the LP2 (SV) fraction from WT and ATP8A1 KO mice (n = 3 independent experiments). b, Hippocampal neurons from WT and ATP8A1 KO mice expressing VAMP7-pH were stimulated at 5, 10, 25 and 50 Hz for 60-150 s in the presence of bafilomycin and the response normalized to NH4Cl (n = 11/3). c, The initial rate of fluorescence increase per action potential in WT and ATP8A1 KO neurons (P = 0.99, 10 Hz; 0.0057, 25 Hz; 2.22 × 10−4, 50 Hz). n = 11/3. d, The recycling pool size in WT and ATP8A1 KO neurons (P = 0.99, 10 Hz; 1.82 × 10−6, 25 Hz; 7.20 × 10−9, 50 Hz). n = 11/3. e, Hippocampal neurons from ATP8A1 KO mice infected with virus encoding WT or E191Q ATP8A1 (with IRES-mCherry) were lysed with SDS-sample buffer and blotted with ATP8A1 antibody. The expression level (right) was normalized to β-action (n = 3 independent experiments. P = 0.3982). n = x/y where x is the number of fields examined and y the number of independent experiments. The data indicate mean ±  s.e.m. for individual coverslips containing 50-100 boutons each. **, P < 0.01; ***, P < 0.001; ****, P < 0.0001 by two-way ANOVA (c,d) with Tukey’s multiple comparisons test or unpaired two-tailed t-test (e).

Source data

Extended Data Fig. 9 High salt releases endogenous synapsins from SVs.

a, WT and mocha neurons expressing GFP-synapsin 1 were stimulated at 5 and 25 Hz and the fluorescence response over boutons normalized to the baseline before stimulation. Analysis (right) shows the time constants of dispersion during stimulation and of reclustering after stimulation (n = 15 coverslips for 5 Hz and 12 coverslips for 25 Hz from three independent experiments. P = 6 × 10−5). b, LP2 (SVs) were treated with high salt (0.3 M glycine, 0.2 M NaCl) for 10-90 min, the membranes sedimented and then immunoblotted for synapsins and VGLUT2 (n = 3 independent experiments). c, Quantitative western analysis shows that LS2 does not contain the membrane protein VGLUT2 (n = 3 independent experiments). d, LS2 fraction contains synapsins. Scatterplot shows the amount of synapsin 1 and 2 in WT and ATP8A1 KO mice (P = 0.16, synapsin 1; 0.016, synapsin 2). n = 3 independent experiments. e, SV fractions (LP2) from WT and ATP8A1 KO mice were treated with high salt buffer to remove endogenous synapsins and then incubated with presynaptic cytosol (LS2) from WT animals in the presence or absence of 0.2 mM ATP. After incubation, the SVs were sedimented and SV-bound synapsin quantified by western analysis. ATP increases synapsin binding to WT but not ATP8A1 KO SVs: synapsin 1 WT (P = 0.0025), ATP8A1 KO (P = 0.1217); synapsin 2 WT (P = 0.0230), ATP8A1 KO (P = 0.9325) and loss of ATP8A1 reduces binding to SVs: synapsin 1, without ATP (P = 4.67 × 10−7); synapsin 1 with ATP (P = 1.41 × 10−8). n = 4 independent experiments The data indicate mean ±  s.e.m. *, P < 0.05; **, P < 0.01; ****, P < 0.0001 by unpaired two-tailed t-test (a,d) or one-way ANOVA (e) with Tukey’s multiple comparison s test.

Source data

Extended Data Fig. 10 The ATP8A1 KO has no effect on mEPSC frequency or amplitude of CA1 pyramidal neurons.

a,b, Representative mEPSCs of CA1 pyramidal neurons from WT and ATP8A1 KO mice (left). Frequency (middle) and amplitude (right) of mEPSCs from WT and ATP8A1 KOs. ns, P = 0.39 for mEPSC frequency; P = 0.84 for mEPSC amplitude by unpaired two-tailed t-test. The data indicate mean ±  s.e.m. n = 6 slices from three animals.

Source data

Supplementary information

Reporting Summary

Supplementary Table 1

Proteomics data

Supplementary Table 2

Sequences of primers

Supplementary Code 1

Code used to analyze the pHluorin imaging

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 3

Unprocessed immunoblots.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 4

Unprocessed immunoblots.

Source Data Fig. 5

Unprocessed immunoblots.

Source Data Fig. 5

Unprocessed immunoblots.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 6

Unprocessed immunoblots.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 6

Unprocessed immunoblots.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 8

Unprocessed immunoblots.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 9

Unprocessed immunoblots.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xu, H., Oses-Prieto, J.A., Khvotchev, M. et al. Adaptor protein AP-3 produces synaptic vesicles that release at high frequency by recruiting phospholipid flippase ATP8A1. Nat Neurosci 26, 1685–1700 (2023). https://doi.org/10.1038/s41593-023-01434-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-023-01434-0

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing