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.

  • Technical Report
  • Published:

Optogenetic acidification of synaptic vesicles and lysosomes

Abstract

Acidification is required for the function of many intracellular organelles, but methods to acutely manipulate their intraluminal pH have not been available. Here we present a targeting strategy to selectively express the light-driven proton pump Arch3 on synaptic vesicles. Our new tool, pHoenix, can functionally replace endogenous proton pumps, enabling optogenetic control of vesicular acidification and neurotransmitter accumulation. Under physiological conditions, glutamatergic vesicles are nearly full, as additional vesicle acidification with pHoenix only slightly increased the quantal size. By contrast, we found that incompletely filled vesicles exhibited a lower release probability than full vesicles, suggesting preferential exocytosis of vesicles with high transmitter content. Our subcellular targeting approach can be transferred to other organelles, as demonstrated for a pHoenix variant that allows light-activated acidification of lysosomes.

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

Figure 1: Design and localization of the light-driven vesicular proton pump pHoenix.
Figure 2: pHoenix can substitute for V-ATPase function.
Figure 3: Activation of pHoenix in untreated neurons increases the quantal size.
Figure 4: Neurons with residual glutamate release after Baf-treatment have smaller mEPSCs and a high paired-pulse ratio.
Figure 5: Partially filled vesicles have a lower release probability.
Figure 6: Optogenetic acidification of lysosomes.

Similar content being viewed by others

Accession codes

Primary accessions

NCBI Reference Sequence

References

  1. Blakely, R.D. & Edwards, R.H. Vesicular and plasma membrane transporters for neurotransmitters. Cold Spring Harb. Perspect. Biol. 4, 1–24 (2012).

    Article  CAS  Google Scholar 

  2. Omote, H., Miyaji, T., Juge, N. & Moriyama, Y. Vesicular neurotransmitter transporter: bioenergetics and regulation of glutamate transport. Biochemistry 50, 5558–5565 (2011).

    Article  CAS  PubMed  Google Scholar 

  3. Mindell, J.A. Lysosomal acidification mechanisms. Annu. Rev. Physiol. 74, 69–86 (2012).

    Article  CAS  PubMed  Google Scholar 

  4. Forgac, M. Vacuolar ATPases: rotary proton pumps in physiology and pathophysiology. Nat. Rev. Mol. Cell Biol. 8, 917–929 (2007).

    Article  CAS  PubMed  Google Scholar 

  5. Boyden, E.S., Zhang, F., Bamberg, E., Nagel, G. & Deisseroth, K. Millisecond-timescale, genetically targeted optical control of neural activity. Nat. Neurosci. 8, 1263–1268 (2005).

    Article  CAS  PubMed  Google Scholar 

  6. Han, X. & Boyden, E.S. Multiple-color optical activation, silencing, and desynchronization of neural activity, with single-spike temporal resolution. PLoS ONE 2, e299 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Chow, B.Y. et al. High-performance genetically targetable optical neural silencing by light-driven proton pumps. Nature 463, 98–102 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Yizhar, O., Fenno, L.E., Davidson, T.J., Mogri, M. & Deisseroth, K. Optogenetics in neural systems. Neuron 71, 9–34 (2011).

    Article  CAS  PubMed  Google Scholar 

  9. Gradinaru, V. et al. Targeting and readout strategies for fast optical neural control in vitro and in vivo. J. Neurosci. 27, 14231–14238 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Greenberg, K.P., Pham, A. & Werblin, F.S. Differential targeting of optical neuromodulators to ganglion cell soma and dendrites allows dynamic control of center-surround antagonism. Neuron 69, 713–720 (2011).

    Article  CAS  PubMed  Google Scholar 

  11. Grubb, M.S. & Burrone, J. Channelrhodopsin-2 localised to the axon initial segment. PLoS ONE 5, e13761 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. Wu, C., Ivanova, E., Cui, J., Lu, Q. & Pan, Z.H. Action potential generation at an axon initial segment-like process in the axonless retinal AII amacrine cell. J. Neurosci. 31, 14654–14659 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Ihara, K. et al. Evolution of the archaeal rhodopsins: evolution rate changes by gene duplication and functional differentiation. J. Mol. Biol. 285, 163–174 (1999).

    Article  CAS  PubMed  Google Scholar 

  14. 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 

  15. Kleinlogel, S. et al. A gene-fusion strategy for stoichiometric and co-localized expression of light-gated membrane proteins. Nat. Methods 8, 1083–1088 (2011).

    Article  CAS  PubMed  Google Scholar 

  16. Takamori, S. VGLUTs: 'exciting' times for glutamatergic research? Neurosci. Res. 55, 343–351 (2006).

    Article  CAS  PubMed  Google Scholar 

  17. Regehr, W.G. Short-term presynaptic plasticity. Cold Spring Harb. Perspect. Biol. 4, a005702 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Wang, L., Tu, P., Bonet, L., Aubrey, K.R. & Supplisson, S. Cytosolic transmitter concentration regulates vesicle cycling at hippocampal GABAergic terminals. Neuron 80, 143–158 (2013).

    Article  CAS  PubMed  Google Scholar 

  19. Herman, M.A., Ackermann, F., Trimbuch, T. & Rosenmund, C. Vesicular glutamate transporter expression level affects synaptic vesicle release probability at hippocampal synapses in culture. J. Neurosci. 34, 11781–11791 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Cousin, M.A. & Nicholls, D.G. Synaptic vesicle recycling in cultured cerebellar granule cells: role of vesicular acidification and refilling. J. Neurochem. 69, 1927–1935 (1997).

    Article  CAS  PubMed  Google Scholar 

  21. Zhou, Q., Petersen, C.C. & Nicoll, R.A. Effects of reduced vesicular filling on synaptic transmission in rat hippocampal neurones. J. Physiol. (Lond.) 525, 195–206 (2000).

    Article  CAS  Google Scholar 

  22. Heine, M. et al. Surface mobility of postsynaptic AMPARs tunes synaptic transmission. Science 320, 201–205 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Liu, G., Choi, S. & Tsien, R.W. Variability of neurotransmitter concentration and nonsaturation of postsynaptic AMPA receptors at synapses in hippocampal cultures and slices. Neuron 22, 395–409 (1999).

    Article  CAS  PubMed  Google Scholar 

  24. Rosenmund, C. & Stevens, C.F. Definition of the readily releasable pool of vesicles at hippocampal synapses. Neuron 16, 1197–1207 (1996).

    Article  CAS  PubMed  Google Scholar 

  25. Metzelaar, M.J. et al. CD63 antigen. A novel lysosomal membrane glycoprotein, cloned by a screening procedure for intracellular antigens in eukaryotic cells. J. Biol. Chem. 266, 3239–3245 (1991).

    CAS  PubMed  Google Scholar 

  26. Mattis, J. et al. Principles for applying optogenetic tools derived from direct comparative analysis of microbial opsins. Nat. Methods 9, 159–172 (2012).

    Article  CAS  Google Scholar 

  27. Hoffmann, A., Hildebrandt, V., Heberle, J. & Buldt, G. Photoactive mitochondria: in vivo transfer of a light-driven proton pump into the inner mitochondrial membrane of Schizosaccharomyces pombe. Proc. Natl. Acad. Sci. USA 91, 9367–9371 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Maycox, P.R., Deckwerth, T. & Jahn, R. Bacteriorhodopsin drives the glutamate transporter of synaptic vesicles after co-reconstitution. EMBO J. 9, 1465–1469 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Moechars, D. et al. Vesicular glutamate transporter VGLUT2 expression levels control quantal size and neuropathic pain. J. Neurosci. 26, 12055–12066 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Pothos, E.N. et al. Synaptic vesicle transporter expression regulates vesicle phenotype and quantal size. J. Neurosci. 20, 7297–7306 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Wojcik, S.M. et al. An essential role for vesicular glutamate transporter 1 (VGLUT1) in postnatal development and control of quantal size. Proc. Natl. Acad. Sci. USA 101, 7158–7163 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Palmer, M.J., Hull, C., Vigh, J. & von Gersdorff, H. Synaptic cleft acidification and modulation of short-term depression by exocytosed protons in retinal bipolar cells. J. Neurosci. 23, 11332–11341 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Williams, J. How does a vesicle know it is full? Neuron 18, 683–686 (1997).

    Article  CAS  PubMed  Google Scholar 

  34. Siksou, L. et al. A role for vesicular glutamate transporter 1 in synaptic vesicle clustering and mobility. Eur. J. Neurosci. 37, 1631–1642 (2013).

    Article  PubMed  Google Scholar 

  35. Budzinski, K.L. et al. Large structural change in isolated synaptic vesicles upon loading with neurotransmitter. Biophys. J. 97, 2577–2584 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Daniels, R.W. et al. Increased expression of the Drosophila vesicular glutamate transporter leads to excess glutamate release and a compensatory decrease in quantal content. J. Neurosci. 24, 10466–10474 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Nelson, H. & Nelson, N. Disruption of genes encoding subunits of yeast vacuolar H+-ATPase causes conditional lethality. Proc. Natl. Acad. Sci. USA 87, 3503–3507 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Davies, S.A. et al. Analysis and inactivation of vha55, the gene encoding the vacuolar ATPase B-subunit in Drosophila melanogaster reveals a larval lethal phenotype. J. Biol. Chem. 271, 30677–30684 (1996).

    Article  CAS  PubMed  Google Scholar 

  39. Inoue, H., Noumi, T., Nagata, M., Murakami, H. & Kanazawa, H. Targeted disruption of the gene encoding the proteolipid subunit of mouse vacuolar H+-ATPase leads to early embryonic lethality. Biochim. Biophys. Acta 1413, 130–138 (1999).

    Article  CAS  PubMed  Google Scholar 

  40. Poëa-Guyon, S. et al. The V-ATPase membrane domain is a sensor of granular pH that controls the exocytotic machinery. J. Cell Biol. 203, 283–298 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Di Giovanni, J. et al. V-ATPase membrane sector associates with synaptobrevin to modulate neurotransmitter release. Neuron 67, 268–279 (2010).

    Article  CAS  PubMed  Google Scholar 

  42. Hiesinger, P.R. et al. The v-ATPase V0 subunit a1 is required for a late step in synaptic vesicle exocytosis in Drosophila. Cell 121, 607–620 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Kawasaki-Nishi, S., Nishi, T. & Forgac, M. Arg-735 of the 100-kDa subunit a of the yeast V-ATPase is essential for proton translocation. Proc. Natl. Acad. Sci. USA 98, 12397–12402 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Kato, H.E. et al. Crystal structure of the channelrhodopsin light-gated cation channel. Nature 482, 369–374 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Müller, M., Bamann, C., Bamberg, E. & Kuhlbrandt, W. Projection structure of channelrhodopsin-2 at 6 A resolution by electron crystallography. J. Mol. Biol. 414, 86–95 (2011).

    Article  PubMed  CAS  Google Scholar 

  46. Yoshimura, K. & Kouyama, T. Structural role of bacterioruberin in the trimeric structure of archaerhodopsin-2. J. Mol. Biol. 375, 1267–1281 (2008).

    Article  CAS  PubMed  Google Scholar 

  47. Pungercar, J.R. et al. Autocatalytic processing of procathepsin B is triggered by proenzyme activity. FEBS J. 276, 660–668 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Weisz, O.A. Organelle acidification and disease. Traffic 4, 57–64 (2003).

    Article  CAS  PubMed  Google Scholar 

  49. Jiang, L.W., Maher, V.M., McCormick, J.J. & Schindler, M. Alkalinization of the lysosomes is correlated with ras transformation of murine and human fibroblasts. J. Biol. Chem. 265, 4775–4777 (1990).

    CAS  PubMed  Google Scholar 

  50. Dehay, B. et al. Loss of P-type ATPase ATP13A2/PARK9 function induces general lysosomal deficiency and leads to Parkinson disease neurodegeneration. Proc. Natl. Acad. Sci. USA 109, 9611–9616 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Royle, S.J., Granseth, B., Odermatt, B., Derevier, A. & Lagnado, L. Imaging pHluorin-based probes at hippocampal synapses. Methods Mol. Biol. 457, 293–303 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Shull, G.E. cDNA cloning of the beta-subunit of the rat gastric H,K-ATPase. J. Biol. Chem. 265, 12123–12126 (1990).

    CAS  PubMed  Google Scholar 

  53. Zhu, Y., Xu, J. & Heinemann, S.F. Two pathways of synaptic vesicle retrieval revealed by single-vesicle imaging. Neuron 61, 397–411 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Shcherbo, D. et al. Far-red fluorescent tags for protein imaging in living tissues. Biochem. J. 418, 567–574 (2009).

    Article  CAS  PubMed  Google Scholar 

  55. Lois, C., Hong, E.J., Pease, S., Brown, E.J. & Baltimore, D. Germline transmission and tissue-specific expression of transgenes delivered by lentiviral vectors. Science 295, 868–872 (2002).

    Article  CAS  PubMed  Google Scholar 

  56. Lock, M. et al. Rapid, simple, and versatile manufacturing of recombinant adeno-associated viral vectors at scale. Hum. Gene Ther. 21, 1259–1271 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Bekkers, J.M. & Stevens, C.F. Excitatory and inhibitory autaptic currents in isolated hippocampal neurons maintained in cell culture. Proc. Natl. Acad. Sci. USA 88, 7834–7838 (1991).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Rost, B.R. et al. Autaptic cultures of single hippocampal granule cells of mice and rats. Eur. J. Neurosci. 32, 939–947 (2010).

    Article  PubMed  Google Scholar 

  59. Edelstein, A., Amodaj, N., Hoover, K., Vale, R. & Stuurman, N. Computer control of microscopes using microManager. Curr. Protoc. Mol. Biol. Ch. 14, Unit 14.20 (2010).

  60. Sankaranarayanan, S., De Angelis, D., Rothman, J.E. & Ryan, T.A. The use of pHluorins for optical measurements of presynaptic activity. Biophys. J. 79, 2199–2208 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Clements, J.D. & Bekkers, J.M. Detection of spontaneous synaptic events with an optimally scaled template. Biophys. J. 73, 220–229 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Carpenter, A.E. et al. CellProfiler: image analysis software for identifying and quantifying cell phenotypes. Genome Biol. 7, R100 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

We thank K. Rösler, B. Brokowski, A. Felies, B. Söhl-Kielczynski, C. Schweynoch, J. Pustogowa and A. Schönherr for excellent technical support, the viral core facility of the Charité for virus production, and M. Camacho-Perez, S. Watanabe and A. Liebkowsky for critical comments on the manuscript. The work was supported by the German Research Council grants Cluster of Excellence Neurocure (EXC 257), Collaborative Research Center SFB958 (D.S., C.R.), Collaborative Research Center SFB1078 (F.S., P.H.), and the Integrative Research Institute for the Life Sciences (D.S., P.H., C.R.), a European Research Council grant (249939 SYNVGLUT; C.R.), a Grass Foundation Fellowship (B.R.R.) and the Louis-Jeantet Foundation (F.S., P.H.).

Author information

Authors and Affiliations

Authors

Contributions

B.R.R. and F.S. developed the concepts for the pHoenix constructs. F.S. performed the molecular biology. B.R.R., F.S., M.K.G., C.W., C.G.B., A.B. and T.R. performed the experiments and analyzed the data. All authors designed the experiments and discussed the results. B.R.R. and F.S. prepared the manuscript, and all authors contributed to editing the paper.

Corresponding author

Correspondence to Benjamin R Rost.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Molecular design of pHoenix and quantification of surface-resident molecules.

(a) Gene-fusion strategy to generate pHoenix, including the amino acid positions of fusion sites and peptide sequences of linker segments (orange). (b) Live-cell images of a pHoenix expressing neuron. The pHluorin signal was acquired in the presence of 50 mM NH4Cl. Scale bar, 50 µm. (c) Heat-colored pHluorin images (averages of 3), acquired during application of pH 5.5 extracellular solution (1), standard extracellular solution (2), and extracellular solution with 50 mM NH4Cl (3). Scale bar, 25 µm. (d) Time-course of pHluorin fluorescence intensities, excited with 100 ms light flashes of a 490-nm LED. Images were acquired at 2 Hz. Signals are baseline corrected. The surface fraction was determined as ratio of fluorescence quenched by the pH 5.5 application (1) to the total range of fluorescence from the pH 5.5 and NH4Cl applications (1 + 3).

Supplementary Figure 2 Expression of pHoenix does not affect synaptic transmitter release.

(a) APs triggered EPSCs of 2.766 ± 0.3 nA in control and 3.3 ± 0.5 nA in pHoenix-expressing neurons. (P = 0.6, U = 681). (b) The size of the readily-releasable pool (RRP) was 0.4 ± 0.05 nC in control and 0.49 ± 0.08 nC in pHoenix-expressing cells (P = 0.9, U = 726). (c) Vesicular release probability (Pvr) was similar for both groups (Pvr(control) = 4.8 ± 0.5 %, Pvr(pHoenix) = 4.9 ± 0.4 %; P = 0.8, U = 715). For (a) – (c): n(control) = 40, n(pHoenix) = 46, N = 4. (d) mEPSC frequency and (e) mEPSC amplitude were not different in both groups: Control: 6.6 ± 1 Hz, 22 ± 1.4 pA (n = 36, N = 4); pHoenix: 5.2 ± 0.8 Hz, 25.5 ± 1.6 pA (n = 43, N = 4); P(frequency) = 0.5, U = 589; P(amplitude) = 0.06, U = 583. (f) Short-term plasticity assessed by train stimulation (50 APs at 10 Hz) was not affected by expression of pHoenix (Control: n = 31, N = 4; pHoenix: n = 40, N = 4). Statistical significance assessed by two-tailed Mann-Whitney U test. Norm., normalized; ns, not significant.

Supplementary Figure 3 Bleaching does not account for the rapid quenching of pHluorin molecules after activation of pHoenix.

(a) Heat-colored images averaged from three photomicrographs of sypHy-expressing neurons 15 s before and after 15 s of illumination with yellow light. Note the high intensity of pHluorin signals in Baf-treated cells compared to untreated cells, indicating a higher vesicular pH after incubation in Baf. Scale bar, 25 µm. (b) Time-course of normalized pHluorin signals in sypHy- or pHoenix-expressing cells. Blue ticks indicate 100 ms excitation of pHluorin with 490 nm light, and numbers indicate acquisition time-points of the example images in (a). Synaptic fluorescence of sypHy-expressing control neurons was stable in untreated cells (black, n = 6, N = 2), and decreased linearly in Baf-treated cells (blue, n = 8, N = 2). In contrast, activation of the proton pump in pHoenix-expressing neurons caused rapid quenching of pHluorin signals (green, n = 15, N = 2; same data as Fig. 2d). Norm., normalized.

Supplementary Figure 4 Application of pHoenix in hippocampal slice cultures.

(a) Schematic illustrating injection of AAV particles into area CA3. (b) pHluorin signals in the pyramidal layer of CA3 in a pHoenix-expressing slice after Baf-treatment for 16 h. Scale bar, 25 µm. (c) Example traces at indicated time points during illumination with 587-nm light, and time plot of EPSC recorded from a CA3 pyramidal neuron. Activation of presynaptic pHoenix by illumination causes a gradual recovery of glutamatergic transmission. Scale bars, 10 ms and 200 pA. (d) Summary of 5 cells recorded in Baf-treated slices expressing pHoenix.

Supplementary Figure 5 Reversible acidification and refilling by pHoenix.

(a) Experimental outline: Cells were recorded after > 2 h of incubation in Baf, and experiments were performed in the presence of 0.1 µM Baf. Every 20 s a single image of pHluorin signals was acquired, followed by train (40 APs at 25 Hz) or sparse (2 APs at 25 Hz) electrophysiological stimulation. pHoenix activity was stimulated by 120-s illumination with 580-nm light, followed by a 300-s dark interval and a second 120-s illumination. (b) A single trace of EPSCs evoked by 40 APs at the end of the first illumination period. AP artifacts deleted for display. Below, only the 1st EPSCs of the train are shown. Illumination caused an increase of the EPSC amplitude, while EPSCs decreased during the dark period. EPSCs recovered during the 2nd illumination. (c) Trace illustrates the sparse stimulation paradigm with 2 APs every 20 s. EPSCs increased during the first light stimulation, but did not significantly decrease in the dark interval. Time points of EPSCs in (b) and (c): Traces during illuminations depict recordings at 0, 20 and 110 s of light, traces during dark interval at 0, 120 and 260 s of depletion (no averages). Scale bars, 4 nA, 100 and 10 ms, respectively. (d) pHluorin signals of the same cell as in (b) with train stimulation. Note the increase of fluorescence at the end of the dark interval due to ongoing exocytosis and lack of reacidification. (e) pHluorin signals of the same cell as in (c), with sparse stimulation. Images in (d) and (e) are averaged from 3 frames, and processed with rolling ball background subtraction for display. Numbers (1–4) indicate acquisition time-points of the example images. Scale bar, 10 µm. (f) Average time course of the 1st EPSC’s amplitudes during train (open symbols) or sparse (closed symbols) stimulation. Both groups: n = 8; N = 3. (g) Time course of pHluorin fluorescence for both stimulation paradigms from the same cells as in (f). Norm., normalized.

Supplementary Figure 6 Vesicular transmitter uptake in Baf-treated neurons is restored by activation of pHoenix, but not Arch3-eGFP, and reaches fill states comparable to those of untreated neurons.

(a) Single EPSC traces and time-plot of EPSCs recorded from a Baf-treated autaptic neuron expressing Arch3-EGFP (gray) or pHoenix (green). Diagram shows averages of EPSCs 30 s after illumination. Transmission was rescued in pHoenix- (n = 42; N = 7), but not Arch3-expressing neurons (n = 10, N = 2; P < 0.0001, Mann-Whitney U test, U = 0). Scale bar, 10 ms, 0.5 nA. (b) EPSCs rescued by pHoenix activation in Baf-treated neurons are not significantly different from the EPSCs in untreated neurons expressing pHoenix after illumination (untreated EPSC: 4.0 ± 1.3 nA before, 4.2 ± 1.4 nA after illumination, n = 20, N = 3; Baf-treated EPSC: 0.1 ± 0.02 nA before, 3.5 ± 0.9 nA after illumination, n = 21, N = 3). Multiplicity adjusted P-values for EPSC amplitudes of untreated vs. treated, pre-light: P = 0.02, post-light: P = 0.87. Traces are average of 6 sweeps before and 30 s after illumination. Scale bar, 1 nA, 10 ms. (c) Traces of mEPSCs recorded from an untreated and a Baf-treated neuron before and after illumination. Scale bars, 50 pA, 200 ms. In Baf-treated cells, mEPSCs were small and occurred at a very low frequency (14.5 ± 1.7 pA; at 0.2 ± 0.05 Hz, n = 19, N = 3, 3 cells displayed no mEPSCs before light). Illumination restored both mEPSC amplitude and frequency (to 28 ± 2 pA, at 5.2 ± 1 Hz), which were not significantly different from the untreated group (22.5 ± 1.9 pA, at 7.6 ± 1.6 Hz before illumination; 26.7 ± 2.3 pA, 7 ± 1.6 Hz after illumination, n = 16, N = 3). Events were detected in a 1 min period before and after illumination. Multiplicity adjusted P-values for mEPSC amplitudes, untreated vs. treated, pre-light: P = 0.015, post-light: P = 0.88; and for mEPSC frequencies, untreated vs. treated, pre-light: P < 0.0001, post-light: P = 0.49 (all tests: Repeated-measures two-way ANOVA, with Sidak's multiple comparisons test). ns, not significant.

Supplementary Figure 7 Pharmacological inhibition of AMPA receptor desensitization does not affect the paired-pulse behavior of residual and rescued EPSCs.

(a) Averages of paired EPSCs of a Baf-treated neuron recorded in the presence of 100 µM cyclothiazide (CTZ) before (Baseline) and after (Full) rescue by pHoenix. AP interval was 40 ms. CTZ slowed EPSC decay kinetics and caused summation of the signals during the second EPSC. Horizontal lines indicate the range of EPSC amplitudes. Scale bars, baseline: 50 ms, 200 pA; full: 50 ms, 1 nA. (b) Recovery of transmission by pHoenix activation in CTZ using low intensity light: EPSC amplitudes increased from 0.55 ± 0.1 nA to 2.53 ± 0.5 nA, with τ = 86 s (P < 0.0001, Wilcoxon signed-rank test, W = 435; n = 29, N = 5). (c) Paired EPSCs scaled to the amplitude of the first EPSC. Scale bar 50 ms. (d) Paired-pulse ratios decreased from 1.2 ± 0.05 for residual EPSCs to 0.96 ± 0.03 for recovered EPSCs after the activation of pHoenix, with τ = 47 s (P < 0.0001, two-tailed paired t-test, t28 = 6). Green curves represent monoexponential association or decay fitted to the data.

Supplementary Figure 8 Vesicular acidification and EPSC amplitudes monitored in parallel during the interval recovery protocol.

During the first illumination period activating pHoenix, pHluorin signals (open circles, normalized to baseline) were rapidly quenched. EPSCs (black filled circles) partially recovered and remained stable during the following dark period, while fluorescence slightly increased due to ongoing exocytosis. EPSCs strongly increased during the second, 90 s illumination period, while pHluorin signals further decreased, indicating that constant proton pumping activity is required for glutamate uptake into synaptic vesicles (n = 9, N = 2). Norm., normalized.

Supplementary Figure 9 Colocalization analysis of lyso-pHoenix and different organelle marker proteins.

HeLa cells transfected with lyso-pHoenix were fixed and stained with antibodies against different markers for major cellular organelles, combined with an anti-GFP staining to enhance the pHluorin signal of the lyso-pHoenix construct. Lyso-pHoenix showed strong colabelling with a staining for LAMP-2 (lysosome-associated membrane protein 2), but no overlap with stainings for the cis-Golgi (GM130, Golgi-Matrix protein 130 kD), mitochondria (Hsp60, 60 kDa heat shock protein), the endoplasmic reticulum (PDI, protein disulphide isomerase) or the early endosome (small GTPase Rab5, Ras-associated protein 5). Scale bar, 10 µm.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rost, B., Schneider, F., Grauel, M. et al. Optogenetic acidification of synaptic vesicles and lysosomes. Nat Neurosci 18, 1845–1852 (2015). https://doi.org/10.1038/nn.4161

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4161

This article is cited by

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