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:

Structural basis of pH-dependent activation in a CLC transporter

An Author Correction to this article was published on 14 February 2024

This article has been updated

Abstract

CLCs are dimeric chloride channels and anion/proton exchangers that regulate processes such as muscle contraction and endo-lysosome acidification. Common gating controls their activity; its closure simultaneously silences both protomers, and its opening allows them to independently transport ions. Mutations affecting common gating in human CLCs cause dominant genetic disorders. The structural rearrangements underlying common gating are unknown. Here, using single-particle cryo-electron microscopy, we show that the prototypical Escherichia coli CLC-ec1 undergoes large-scale rearrangements in activating conditions. The slow, pH-dependent remodeling of the dimer interface leads to the concerted opening of the intracellular H+ pathways and is required for transport. The more frequent formation of short water wires in the open H+ pathway enables Cl pore openings. Mutations at disease-causing sites favor CLC-ec1 activation and accelerate common gate opening in the human CLC-7 exchanger. We suggest that the pH activation mechanism of CLC-ec1 is related to the common gating of CLC-7.

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: pH- and Cl-dependent conformational rearrangements of CLC-ec1.
Fig. 2: Opening of the H+ pathway is required for transport by CLC-ec1.
Fig. 3: smFRET-TIRF microscopy of interprotomer dynamics in a membrane environment.
Fig. 4: Substrate accessibilities of the H+ and Cl pathways increase in Turn and Twist.
Fig. 5: Mechanism of pH-dependent activation of CLC-ec1.
Fig. 6: Activation mechanism of CLC-ec1.

Similar content being viewed by others

Data availability

All constructs are available on request to A.A. or O.B. All models and associated cryo-EM maps have been deposited into the Electron Microscopy Data Bank (EMDB) and the Protein Data Bank (PDB) under the following accession codes CLC-ec1 pH 7.5 0 mM Cl Swap (PDB: 7RNX; EMD-24582); CLC-ec1 pH 7.5 100 mM Cl Swap (PDB: 7RO0; EMD-24584); CLC-ec1 pH 4.5 0 mM Cl Swap (PDB: 7RP6; EMD-24668); CLC-ec1 pH 4.5 0 mM Cl Turn (PDB:7RSB; EMD-24583); CLC-ec1 pH 4.5 100 mM Cl Swap (PDB: 7N8P; EMD-24241); CLC-ec1 pH 4.5 100 mM Cl Turn (PDB: 7RP5; EMD-24612); CLC-ec1 pH 4.5 100 mM Cl Twist (PDB: 7N0W;EMD-24263); R230C L249C pH 4.5 100 mM Cl Swap (PDB: 8GA1; 29884); R230C L249C pH 4.5 100 mM Cl Turn (PDB: 8GA3; EMD-29885); L25C A450C pH 4.5 100 mM Cl Intermediate (PDB:8GA5; EMD-29890); L25C A450C pH 4.5 100 mM Cl Twist (PDB: 8GAH; EMD-29899); E202Y pH 4.5 100 mM Cl (PDB: 8GA0; EMD-29883), which are listed in Supplementary Table 7. The atomic coordinates of the free MD simulation at t = 0 and 1 µs of systems 1–8 listed in Supplementary Table 3 are available at https://doi.org/10.6084/m9.figshare.24777321.v1. Source data are provided with this paper.

Change history

References

  1. Accardi, A. Structure and gating of CLC channels and exchangers. J. Physiol. 593, 4129–4138 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Jentsch, T. J. & Pusch, M. CLC chloride channels and transporters: structure, function, physiology, and disease. Physiol. Rev. 98, 1493–1590 (2018).

    CAS  PubMed  Google Scholar 

  3. Miller, C. Open-state substructure of single chloride channels from Torpedo electroplax. Philos. Trans. R. Soc. Lond. B Biol. Sci. 299, 401–411 (1982).

    CAS  PubMed  Google Scholar 

  4. Ludewig, U., Pusch, M. & Jentsch, T. J. Independent gating of single pores in CLC-0 chloride channels. Biophys. J. 73, 789–797 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Weinreich, F. & Jentsch, T. J. Pores formed by single subunits in mixed dimers of different CLC chloride channels. J. Biol. Chem. 276, 2347–2353 (2001).

    CAS  PubMed  Google Scholar 

  6. Middleton, R. E., Pheasant, D. J. & Miller, C. Homodimeric architecture of a ClC-type chloride ion channel. Nature 383, 337–340 (1996).

    CAS  PubMed  Google Scholar 

  7. Saviane, C., Conti, F. & Pusch, M. The muscle chloride channel ClC-1 has a double-barreled appearance that is differentially affected in dominant and recessive myotonia. J. Gen. Physiol. 113, 457–468 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Jentsch, T. J., Steinmeyer, K. & Schwarz, G. Primary structure of Torpedo marmorata chloride channel isolated by expression cloning in Xenopus oocytes. Nature 348, 510–514 (1990).

    CAS  PubMed  Google Scholar 

  9. Fischer, M., Janssen, A. G. & Fahlke, C. Barttin activates ClC-K channel function by modulating gating. J. Am. Soc. Nephrol. 21, 1281–1289 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Stölting, G., Fischer, M. & Fahlke, C. ClC-1 and ClC-2 form hetero-dimeric channels with novel protopore functions. Pflugers Arch. 466, 2191–2204 (2014).

    PubMed  Google Scholar 

  11. Pusch, M., Ludewig, U., Rehfeldt, A. & Jentsch, T. J. Gating of the voltage-dependent chloride channel CIC-0 by the permeant anion. Nature 373, 527–531 (1995).

    CAS  PubMed  Google Scholar 

  12. Accardi, A. & Pusch, M. Fast and slow gating relaxations in the muscle chloride channel CLC-1. J. Gen. Physiol. 116, 433–444 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Zúñiga, L. et al. The voltage-dependent ClC-2 chloride channel has a dual gating mechanism. J. Physiol. 555, 671–682 (2004).

    PubMed  PubMed Central  Google Scholar 

  14. Zifarelli, G., Pusch, M. & Fong, P. Altered voltage-dependence of slowly activating chloride-proton antiport by late endosomal ClC-6 explains distinct neurological disorders. J. Physiol. 600, 2147–2164 (2022).

    CAS  PubMed  Google Scholar 

  15. Matsuda, J. J. et al. Overexpression of CLC-3 in HEK293T cells yields novel currents that are pH dependent. Am. J. Physiol. Cell Physiol. 294, 251–262 (2008).

    Google Scholar 

  16. Ludwig, C. F., Ullrich, F., Leisle, L., Stauber, T. & Jentsch, T. J. Common gating of both CLC transporter subunits underlies voltage-dependent activation of the 2Cl/1H+ exchanger ClC-7/Ostm1. J. Biol. Chem. 288, 28611–28619 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Leisle, L., Ludwig, C. F., Wagner, F. A., Jentsch, T. J. & Stauber, T. ClC-7 is a slowly voltage-gated 2Cl/1H+-exchanger and requires Ostm1 for transport activity. EMBO J. 30, 2140–2152 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. De Stefano, S., Pusch, M. & Zifarelli, G. A single point mutation reveals gating of the human ClC-5 Cl/H+ antiporter. J. Physiol. 591, 5879–5893 (2013).

    PubMed  PubMed Central  Google Scholar 

  19. Alekov, A. K. & Fahlke, C. Channel-like slippage modes in the human anion/proton exchanger ClC-4. J. Gen. Physiol. 133, 485–496 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Suetterlin, K. et al. Translating genetic and functional data into clinical practice: a series of 223 families with myotonia. Brain 145, 607–620 (2022).

    PubMed  Google Scholar 

  21. Pusch, M. Myotonia caused by mutations in the muscle chloride channel gene CLCN1. Hum. Mutat. 19, 423–434 (2002).

    CAS  PubMed  Google Scholar 

  22. Dupré, N. et al. Clinical, electrophysiologic, and genetic study of non-dystrophic myotonia in French-Canadians. Neuromuscul. Disord. 19, 330–334 (2009).

    PubMed  Google Scholar 

  23. Altamura, C. et al. The analysis of myotonia congenita mutations discloses functional clusters of amino acids within the CBS2 domain and the C-terminal peptide of the ClC-1 channel. Hum. Mutat. 39, 1273–1283 (2018).

    CAS  PubMed  Google Scholar 

  24. Zifarelli, G. The role of the lysosomal Cl/H+ Antiporter ClC-7 in osteopetrosis and neurodegeneration. Cells 11, 366 (2022).

  25. Leisle, L. et al. Divergent Cl and H+ pathways underlie transport coupling and gating in CLC exchangers and channels. eLife 9, e51224 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Accardi, A. et al. Separate ion pathways in a Cl/H+ exchanger. J. Gen. Physiol. 126, 563–570 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Pusch, M., Ludewig, U. & Jentsch, T. J. Temperature dependence of fast and slow gating relaxations of ClC-0 chloride channels. J. Gen. Physiol. 109, 105–116 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Bykova, E. A., Zhang, X. D., Chen, T. Y. & Zheng, J. Large movement in the C terminus of CLC-0 chloride channel during slow gating. Nat. Struct. Mol. Biol. 13, 1115–1119 (2006).

    CAS  PubMed  Google Scholar 

  29. Yu, Y., Tsai, M. F., Yu, W. P. & Chen, T. Y. Modulation of the slow/common gating of CLC channels by intracellular cadmium. J. Gen. Physiol. 146, 495–508 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Abraham, S. J. et al. 13C NMR detects conformational change in the 100-kD membrane transporter ClC-ec1. J. Biomol. NMR 61, 209–226 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Chavan, T. S. et al. A CLC-ec1 mutant reveals global conformational change and suggests a unifying mechanism for the CLC Cl/H+ transport cycle. eLife 9, e53479 (2020).

    PubMed  PubMed Central  Google Scholar 

  32. Khantwal, C. M. et al. Revealing an outward-facing open conformational state in a CLC Cl/H+ exchange transporter. eLife 5, e11189 (2016).

    PubMed  PubMed Central  Google Scholar 

  33. Heath, G. R. et al. Localization atomic force microscopy. Nature 594, 385–390 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Punjani, A., Rubinstein, J. L., Fleet, D. J. & Brubaker, M. A. cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods 14, 290 (2017).

    CAS  PubMed  Google Scholar 

  35. Lim, H. H., Shane, T. & Miller, C. Intracellular proton access in a Cl/H+. antiporter. PLoS Biol. 10, e1001441 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Han, W., Cheng, R. C., Maduke, M. C. & Tajkhorshid, E. Water access points and hydration pathways in CLC H+/Cl transporters. Proc. Natl Acad. Sci. USA 111, 1819–1824 (2014).

    CAS  PubMed  Google Scholar 

  37. Lee, S., Swanson, J. M. & Voth, G. A. Multiscale simulations reveal key aspects of the proton transport mechanism in the ClC-ec1 antiporter. Biophys. J. 110, 1334–1345 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Chadda, R. et al. The dimerization equilibrium of a ClC Cl/H+ antiporter in lipid bilayers. eLife 5, e17438 (2016).

    PubMed  PubMed Central  Google Scholar 

  39. Nguitragool, W. & Miller, C. CLC Cl/H+ transporters constrained by covalent cross-linking. Proc. Natl Acad. Sci. USA 104, 20659–20665 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Robertson, J. L., Kolmakova-Partensky, L. & Miller, C. Design, function and structure of a monomeric ClC transporter. Nature 468, 844–847 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Juette, M. F. et al. Single-molecule imaging of non-equilibrium molecular ensembles on the millisecond timescale. Nat. Methods 13, 341–344 (2016).

    PubMed  PubMed Central  Google Scholar 

  42. Altman, R. B. et al. Enhanced photostability of cyanine fluorophores across the visible spectrum. Nat. Methods 9, 428–429 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Zheng, Q., Jockusch, S., Zhou, Z. & Blanchard, S. C. The contribution of reactive oxygen species to the photobleaching of organic fluorophores. Photochem. Photobiol. 90, 448–454 (2014).

    CAS  PubMed  Google Scholar 

  44. Walden, M. et al. Uncoupling and turnover in a Cl/H+ exchange transporter. J. Gen. Physiol. 129, 317–329 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Accardi, A. & Miller, C. Secondary active transport mediated by a prokaryotic homologue of ClC Cl channels. Nature 427, 803–807 (2004).

    CAS  PubMed  Google Scholar 

  46. Lim, H. H. & Miller, C. Intracellular proton-transfer mutants in a CLC Cl/H+ exchanger. J. Gen. Physiol. 133, 8 (2009).

    Google Scholar 

  47. Feng, L., Campbell, E. B., Hsiung, Y. & MacKinnon, R. Structure of a eukaryotic CLC transporter defines an intermediate state in the transport cycle. Science 330, 635–641 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Zdebik, A. A. et al. Determinants of anion-proton coupling in mammalian endosomal CLC proteins. J. Biol. Chem. 283, 4219–4227 (2008).

    CAS  PubMed  Google Scholar 

  49. Wang, D. & Voth, G. A. Proton transport pathway in the ClC Cl/H+ antiporter. Biophys. J. 97, 121–131 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Koty, P. P., Pegoraro, E. & Hoffman, E. P. Linkage and mutation analysis of Thomsen and Becker myotonia families. Am. J. Hum. Genet. 55, A227.1323 (1994).

    Google Scholar 

  51. Lehmann-Horn, F., Mailänder, V., Heine, R. & George, A. L. Myotonia levior is a chloride channel disorder. Hum. Mol. Genet. 4, 1397–1402 (1995).

    CAS  PubMed  Google Scholar 

  52. Palmer, E. E. et al. Functional and clinical studies reveal pathophysiological complexity of CLCN4-related neurodevelopmental condition. Mol. Psychiatry 28, 668–697 (2023).

    CAS  PubMed  Google Scholar 

  53. Meyer-Kleine, C., Steinmeyer, K., Ricker, K., Jentsch, T. J. & Koch, M. C. Spectrum of mutations in the major human skeletal muscle chloride channel gene CLCN1 leading to myotonia. Am. J. Hum. Genet. 57, 1325–1334 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Hoopes, R. R. et al. Evidence for genetic heterogeneity in Dent’s disease. Kidney Int. 65, 1615–1620 (2004).

    CAS  PubMed  Google Scholar 

  55. Halbritter, J. et al. Fourteen monogenic genes account for 15% of nephrolithiasis/nephrocalcinosis. J. Am. Soc. Nephrol. 26, 543–551 (2015).

    CAS  PubMed  Google Scholar 

  56. Minamikawa, S. et al. Development of ultra-deep targeted RNA sequencing for analyzing X-chromosome inactivation in female Dent disease. J. Hum. Genet. 63, 589–595 (2018).

    CAS  PubMed  Google Scholar 

  57. Tosetto, E. et al. Phenotypic and genetic heterogeneity in Dent’s disease—the results of an Italian collaborative study. Nephrol. Dial. Transpl. 21, 2452–2463 (2006).

    CAS  Google Scholar 

  58. Wang, C. et al. The virulence gene and clinical phenotypes of osteopetrosis in the Chinese population: six novel mutations of the CLCN7 gene in twelve osteopetrosis families. J. Bone Miner. Metab. 30, 338–348 (2012).

    PubMed  Google Scholar 

  59. Li, L., Lv, S.-S., Wang, C., Yue, H. & Zhang, Z.-L. Novel CLCN7 mutations cause autosomal dominant osteopetrosis type II and intermediate autosomal recessive osteopetrosis. Mol. Med. Rep. 19, 5030–5038 (2019).

    CAS  PubMed  Google Scholar 

  60. Zhixuan, Z., Long, C., Jin, H. & Ji, S. Structure of the human CLC-7/Ostm1 complex reveals a novel state. JUSTC 53, 0306-0301-0306-0307 (2023).

    Google Scholar 

  61. Stauber, T. & Jentsch, T. J. Sorting motifs of the endosomal/lysosomal CLC chloride transporters. J. Biol. Chem. 285, 34537–34548 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Dutzler, R., Campbell, E. B., Cadene, M., Chait, B. T. & MacKinnon, R. X-ray structure of a ClC chloride channel at 3.0 Å reveals the molecular basis of anion selectivity. Nature 415, 287–294 (2002).

    CAS  PubMed  Google Scholar 

  63. Dutzler, R., Campbell, E. B. & MacKinnon, R. Gating the selectivity filter in ClC chloride channels. Science 300, 108–112 (2003).

    CAS  PubMed  Google Scholar 

  64. Park, E. & MacKinnon, R. Structure of the CLC-1 chloride channel from Homo sapiens. eLife 7, e36629 (2018).

    PubMed  PubMed Central  Google Scholar 

  65. Park, E., Campbell, E. B. & MacKinnon, R. Structure of a CLC chloride ion channel by cryo-electron microscopy. Nature 541, 500–505 (2017).

    CAS  PubMed  Google Scholar 

  66. Feng, L., Campbell, E. B. & MacKinnon, R. Molecular mechanism of proton transport in CLC Cl/H+ exchange transporters. Proc. Natl Acad. Sci. USA 109, 11699–11704 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Mayes, H. B., Lee, S., White, A. D., Voth, G. A. & Swanson, J. M. J. Multiscale kinetic modeling reveals an ensemble of Cl/H+ exchange pathways in ClC-ec1 antiporter. J. Am. Chem. Soc. 140, 1793–1804 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Basilio, D., Noack, K., Picollo, A. & Accardi, A. Conformational changes required for H+/Cl exchange mediated by a CLC transporter. Nat. Struct. Mol. Biol. 21, 456–463 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Accardi, A. & Pusch, M. Conformational changes in the pore of CLC-0. J. Gen. Physiol. 122, 277–293 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Bell, S. P., Curran, P. K., Choi, S. & Mindell, J. A. Site-directed fluorescence studies of a prokaryotic ClC antiporter. Biochemistry 45, 6773–6782 (2006).

    CAS  PubMed  Google Scholar 

  71. Osteen, J. & Mindell, J. A. Zn2+ inhibition of CLC-4. Biophys. J. 95, 4668–4675 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Elvington, S., Liu, C. & Maduke, M. Substrate-driven conformational changes in ClC-ec1 observed by fluorine NMR. EMBO J. 28, 3090–3102 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Pusch, M. Chloride channelopathies. Pharm. News 8, 45–51 (2001).

    CAS  Google Scholar 

  74. Kubisch, C., Schmidt-Rose, T., Fontaine, B., Bretag, A. H. & Jentsch, T. J. ClC-1 chloride channel mutations in myotonia congenita: variable penetrance of mutations shifting the voltage dependence. Hum. Mol. Genet. 7, 1753–1760 (1998).

    CAS  PubMed  Google Scholar 

  75. Lourdel, S. et al. ClC-5 mutations associated with Dent’s disease: a major role of the dimer interface. Pflugers Arch. 463, 247–256 (2012).

    CAS  PubMed  Google Scholar 

  76. Leray, X. et al. Tonic inhibition of the chloride/proton antiporter ClC-7 by PI(3,5)P2 is crucial for lysosomal pH maintenance. eLife 11, e74136 (2022).

  77. Zhang, S. et al. Molecular insights into the human CLC-7/Ostm1 transporter. Sci. Adv. 6, eabb4747 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Smart, O. S., Neduvelil, J. G., Wang, X., Wallace, B. A. & Sansom, M. S. HOLE: a program for the analysis of the pore dimensions of ion channel structural models. J. Mol. Graph 14, 354–360 (1996).

    CAS  PubMed  Google Scholar 

  79. Schneider, C. A., Rasband, W. S. & Eliceiri, K. W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 9, 671 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Fortea, E. & Accardi, A. A quantitative flux assay for the study of reconstituted Cl channels and transporters. Methods Enzymol. 652, 243–272 (2021).

    CAS  PubMed  Google Scholar 

  81. Russo, C. J. & Passmore, L. A. Electron microscopy: ultrastable gold substrates for electron cryomicroscopy. Science 346, 1377–1380 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Punjani, A., Zhang, H. & Fleet, D. J. Non-uniform refinement: adaptive regularization improves single-particle cryo-EM reconstruction. Nat. Methods 17, 1214–1221 (2020).

    CAS  PubMed  Google Scholar 

  83. Punjani, A. & Fleet, D. J. 3D variability analysis: resolving continuous flexibility and discrete heterogeneity from single particle cryo-EM. J. Struct. Biol. 213, 107702 (2021).

    CAS  PubMed  Google Scholar 

  84. Pettersen, E. F. et al. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004).

    CAS  PubMed  Google Scholar 

  85. Liebschner, D. et al. Macromolecular structure determination using X-rays, neutrons and electrons: recent developments in Phenix. Acta Crystallogr. D Struct. Biol. 75, 861–877 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D Biol. Crystallogr. 60, 2126–2132 (2004).

    PubMed  Google Scholar 

  87. Ciftci, D. et al. FRET-based microscopy assay to measure activity of membrane amino acid transporters with single-transporter resolution. Bio. Protoc. 11, e3970 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Joo, C. & Ha, T. Single-molecule FRET with total internal reflection microscopy. Cold Spring Harb. Protoc. https://doi.org/10.1101/pdb.top072058 (2012).

  89. Akyuz, N. et al. Transport domain unlocking sets the uptake rate of an aspartate transporter. Nature 518, 68–73 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  90. Huysmans, G. H. M., Ciftci, D., Wang, X., Blanchard, S. C. & Boudker, O. The high-energy transition state of the glutamate transporter homologue GltPh. EMBO J. 40, e105415 (2020).

  91. Aitken, C. E., Marshall, R. A. & Puglisi, J. D. An oxygen scavenging system for improvement of dye stability in single-molecule fluorescence experiments. Biophys. J. 94, 1826–1835 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Sigworth, F. J. & Sine, S. M. Data transformations for improved display and fitting of single-channel dwell time histograms. Biophys. J. 52, 1047–1054 (1987).

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Di Zanni, E. et al. Pathobiologic mechanisms of neurodegeneration in osteopetrosis derived from structural and functional analysis of 14 ClC-7 mutants. J. Bone Miner. Res. 36, 531–545 (2021).

    PubMed  Google Scholar 

  94. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph. 14, 33–38 (1996).

    CAS  PubMed  Google Scholar 

  95. Wu, E. L. et al. CHARMM-GUI Membrane Builder toward realistic biological membrane simulations. J. Comput. Chem. 35, 1997–2004 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Best, R. B. et al. Optimization of the Additive CHARMM All-atom protein force field targeting improved sampling of the backbone ϕ, ψ and side-chain χ1 and χ2 dihedral angles. J. Chem. Theory Comput. 8, 3257–3273 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., Impey, R. W. & Klein, M. L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79, 926–935 (1983).

    CAS  Google Scholar 

  98. Abraham, M. J. et al. GROMACS: high performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 1-2, 19–25 (2015).

    Google Scholar 

  99. Tribello, G. A., Bonomi, M., Branduardi, D., Camilloni, C. & Bussi, G. PLUMED 2: new feathers for an old bird. Comput. Phys. Commun. 185, 604–613 (2014).

    CAS  Google Scholar 

  100. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: an Nlog(N) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    CAS  Google Scholar 

  101. Berk, H., Henk, B., Berendsen, H. J. C. & Fraaije, J. G. E. M. LINCS: a linear constraint solver for molecular simulations. J. Comput. Chem. 18, 1463–1472 (1997).

    Google Scholar 

  102. Søndergaard, C. R., Olsson, M. H. M., Rostkowski, M. & Jensen, J. H. Improved treatment of ligands and coupling effects in empirical calculation and rationalization of pKa values. J. Chem. Theory Comput. 7, 2284–2295 (2011).

    PubMed  Google Scholar 

  103. Olsson, M. H. M., Søndergaard, C. R., Rostkowski, M. & Jensen, J. H. PROPKA3: consistent treatment of internal and surface residues in empirical pKa predictions. J. Chem. Theory Comput. 7, 525–537 (2011).

    CAS  PubMed  Google Scholar 

  104. Im, W. & Roux, B. T. Ions and counterions in a biological channel: a molecular dynamics simulation of OmpF Porin from Escherichia coli in an explicit membrane with 1 M KCl aqueous salt solution. J. Mol. Biol. 319, 1177–1197 (2002).

    CAS  PubMed  Google Scholar 

  105. Salari, R., Joseph, T., Lohia, R., Hénin, J. & Brannigan, G. A streamlined, general approach for computing ligand binding free energies and its application to GPCR-bound cholesterol. J. Chem. Theory Comput. 14, 6560–6573 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Alleva, C. et al. Na+-dependent gate dynamics and electrostatic attraction ensure substrate coupling in glutamate transporters. Sci. Adv. 6, eaba9854 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Sugita, Y., Kitao, A. & Okamoto, Y. Multidimensional replica-exchange method for free-energy calculations. J. Chem. Phys. 113, 6042–6051 (2000).

    CAS  Google Scholar 

  108. Lee, S., Liang, R., Voth, G. A. & Swanson, J. M. Computationally efficient multiscale reactive molecular dynamics to describe amino acid deprotonation in proteins. J. Chem. Theory Comput. 12, 879–891 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Domański, J., Sansom, M. S. P., Stansfeld, P. J. & Best, R. B. Atomistic mechanism of transmembrane helix association. PLoS Comput. Biol. 16, e1007919 (2020).

    PubMed  PubMed Central  Google Scholar 

  110. Kumar, S., Rosenberg, J. M., Bouzida, D., Swendsen, R. H. & Kollman, P. A. The weighted histogram analysis method for free-energy calculations on biomolecules. I. The method. J. Comput. Chem. 13, 1011–1021 (1992).

    CAS  Google Scholar 

  111. Friedman, L. J., Chung, J. & Gelles, J. Viewing dynamic assembly of molecular complexes by multi-wavelength single-molecule fluorescence. Biophys. J. 91, 1023–1031 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Friedman, L. J. & Gelles, J. Multi-wavelength single-molecule fluorescence analysis of transcription mechanisms. Methods 86, 27–36 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank members of the Accardi and Boudker labs, C. Nimigean and S. Scheuring for helpful discussions. We thank R. Hite for help with cryo-EM data analysis. We thank S. Blanchard for help with establishing smFRET resources. The work was supported by the National Institutes of Health (NIH) grants R01 GM128420 and R35 GM152012 (to A.A.) and R01GM120260 and R21GM126476 (to J.L.R.), and American Heart Association grant 19PRE34380215 (to D.C.). The electron microcopy data were collected on the Tiran Krios at the Cryo-EM Core Facility at NYU Langone Health, with the assistance of W. Rice and B. Wang (RRID: SCR_019202). Some of this work was performed at the Simons Electron Microscopy Center and National Resource for Automated Molecular Microscopy located at the New York Structural Biology Center, supported by grants from the Simons Foundation (SF349247), NYSTAR and the NIH National Institute of General Medical Sciences (GM103310). The computational resource of this research was provided in part by TACC Stampede2, SDSC Expanse, PSC Bridges2 and NCSA Delta through allocation number MCB200216 from the Extreme Science and Engineering Discovery Environment (XSEDE), which was supported by National Science Foundation grant number no. 1548562, and in part by HPC resources supported by the Scientific Computing Unit at Weill Cornell Medicine.

Author information

Authors and Affiliations

Authors

Contributions

E.F. designed and performed cryo-EM experiments of WT CLC-ec1, collected and carried out initial processing of the CLC-ec1-E202Y dataset, smFRET measurements, patch–clamp recordings of CLC-7 and flux assays of WT and mutant CLC-ec1; S.L. designed and performed cryo-EM experiments and flux assays of Top and Bot crosslinked CLC-ec1, collected the CLC-ec1-E202Y dataset, completed processing and built the atomic model, carried out MD simulations and performed flux assays on mutant CLC-ec1; R.C., P.S., R.M.-K. and J.L.R. designed, carried out and analyzed dimerization experiments; Y.A. contributed to the electrophysiological experiments; M.E.F. contributed to the initial cryo-EM data processing; D.C. and G.H. contributed to the design of smFRET experiments, data acquisition and analysis; and O.B. and A.A. oversaw research and wrote the initial draft of the paper. All authors edited and approved the paper.

Corresponding authors

Correspondence to Olga Boudker or Alessio Accardi.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Structural & Molecular Biology thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available. Primary Handling Editor: Katarzyna Ciazynska and Carolina Perdigoto, in collaboration with the Nature Structural & Molecular Biology team.

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 CryoEM image processing scheme for the dataset at pH 4.5 in 100 mM Cl−.

(a) Representative cryoEM micrographs of CLC-ec1. (b) Representative 2D classes from CryoSPARC. (c) Summary of the image processing procedure. Red frames indicate classes that were selected for further processing. (d–f) Density maps (mesh) for αB, αH, αI, and the H-I loop with the refined models of Swap (D), Turn (E), and Twist (F). (g–j) EM density (mesh) and atomic models of E148 for Swap at pH 7.5 and 100 mM Cl− (G), Swap at pH 7.5 and 0 mM Cl− (H), Turn at pH 4.5 and 0 mM Cl− (I), and Twist and pH 4.5 and 100 mM Cl− (J). Data acquisition and refinement parameters are reported in Tables 1 and 2.

Extended Data Fig. 2 Validation of maps and models for WT CLC-ec1.

Fourier shell correlation (FSC) curves for the masked (tight mask from CryoSPARC) and unmasked density maps (left), FSC curves from model−map cross-validation (middle), and density maps colored by local resolution estimated in CryoSPARC (right) of maps obtained in the following conditions: pH 4.5 and 100 mM Cl− (A), pH 4.5 and 0 mM Cl− (B), pH 7.5 and 0 mM Cl− (C), pH 7.5 and 100 mM Cl− (D). Data acquisition and refinement parameters are reported in Table 1.

Source data

Extended Data Fig. 3 Structural comparisons between different CLC-ec1 conformations.

(a, b) Per residue Cα r.m.s.d. for CLC-ec1 in Swap at pH 7.5 and 100 mM Cl− relative to the crystal structure (PDB: 1OTS) (A) and for Turn at pH 4.5 and 100 mM Cl− relative to E113Q/E148Q/E203Q (QQQ, PDB: 6V2J) (B). (c–e) Close-up view of the rearrangements of the H-I loops in Swap (C), Turn (D), and Twist (E). The top panels are viewed from the plane of the membrane, and the bottom panels are viewed from the intracellular side. One protomer is colored gray, and the other is colored red (Swap), wheat (Turn), and Blue (Twist). Helices αH and αI are shown in cartoon representation, and residues Q207, F208, R209, Y210, T211, and L212 are shown in stick CPK representation. (f) Quantification of the surface area per protomer buried at the interface of the soluble (gray bars) and TM (black bars) domains in Swap, Turn, and Twist. (g, h) Cartoon representations of single protomers viewed from the dimer interface and colored by changes in per-residue contact areas in Turn (G) and Twist (H) relative to Swap. (i) Diameter of the Cl− permeation pathway visualized using HOLE 1 in cryoEM structures. The position of the external gate is defined as the midpoint of the backbone atoms of residues 147, 148, 356, and 357. Residues S107, E148, I356, and Y445 are shown as sticks. Blue indicates regions with R > 2.3 Å, green 2.3 Å>R > 1.15 Å, and red R < 1.15 Å.

Source data

Extended Data Fig. 4 Characterization of Top and Bot crosslinks.

(a, b) SDS PAGE showing the molecular mass of TCEP-treated (right) or untreated (left) CLC-ec1 WT, Top (R230C/L249C) and Bot (L25C/A450C) crosslinks, L25C and A450C single mutants, and the Top+Bot quadruple mutant. Arrows denoted by D and M indicate the position of the dimer and monomer, respectively. Note that L25C on helix αA spontaneously forms dimers, reflecting crosslinks between the two dynamic αA helices. This results in two dimer bands for the Bot and Top+Bot crosslinked constructs, where αA of one protomer is crosslinked either to αA or αR of the sister protomer. SDS PAGE results were consistent in three repeats from independently obtained proteo-liposome samples of WT and all mutants. (c) Time courses of Cl− efflux mediated by TCEP-treated (right) or untreated (left) proteoliposomes reconstituted with CLC-ec1 WT (black/blue), Bot (pink/dark green), L25C (light green/gray), Top (orange/brown), A450C (purple/yellow) and Top+Bot (red/cyan). Traces for WT and Bot ±TCEP are the same as those shown in Fig. 2. (d) Intracellular view of cryoEM density maps of CLC-ec1 Bot crosslink in Intermediate (left panel) and Twist (right panel) conformations imaged at pH 4.5 in 100 mM Cl−. One subunit is colored gray, and the other is pink (Intermediate) or yellow (Twist). (e) CryoEM density (mesh) contoured at 5.1 σ and atomic model of the crosslink between A450C and L25C in Bot Intermediate. (f) Intracellular view of cryoEM density maps of CLC-ec1 Top crosslink in Swap (left panel) and Turn (right panel) conformations imaged at pH 4.5 in 100 mM Cl−. One subunit is colored gray, and the other is green (Swap) or aquamarine (Turn). (g) CryoEM density (mesh) contoured at 6.0 σ and atomic model of the crosslink between A450C and L25C in Top Swap.

Source data

Extended Data Fig. 5 Validation of maps and models for CLC-ec1 mutants.

(a–c) FSC curves for the masked (tight mask from CryoSPARC) and unmasked density maps (left), FSC curves from model−map cross-validation (middle), and density maps colored by local resolution estimated in CryoSPARC (right) of maps obtained for CLC-ec1 (A) L25C/A450C Bot Intermediate and Twist, (B) R230C/L249C Top Swap and Turn, (C) E202Y. All data were collected at pH 4.5 and 100 mM Cl−. Data acquisition and refinement parameters are reported in Table 2.

Source data

Extended Data Fig. 6 Characterization of CLC-ec1 mutants used in smFRET experiments and the dynamics of the dimer interface.

(a) Representative normalized time course of chloride efflux from liposomes reconstituted at 0.2 μg protein/mg lipid with CLC-ec1 WT (black), Q24C (orange) and Q24C labelled with maleimide activated LD555p and LD650 fluorophores in the same molar ratio used in smFRET TIRF imaging (CLC-ec1:LD555p:LD650 = 1:2:2.5) (green). Traces were normalized to the total chloride content of the liposomes determined following the addition of detergent2. (b) The initial velocity of Cl efflux from proteoliposomes containing WT CLC-ec1 (red), CLC-ec1 Q24C (orange), and Q24C-labelled (green) All values are means ± S.D. The number of repeats of independent experiments and preparations of proteo-liposome samples for all constructs are reported in Supplementary Table 1. (c) The experimental design to monitor dynamics of the TM dimer interface tracking the movement of E385C labelled with donor and acceptor fluorophores. CLC-ec1 dimer is shown as cartoon and colored as in Fig. 1. The labeling site is shown as a black sphere, and inter-dye distances are shown on the structures. Schematic representations of Swap, Turn, and Twist are on the left. (d-g) Population contour plots of labelled E385C at pH 7.5 in 0 mM Cl (D), pH 7.5 in 100 mM Cl (E), pH 4.5 in 0 mM Cl (F), and pH 4.5 in 100 mM Cl (G). N, the number of molecules used to construct the plots, is shown on each panel. Cumulative FRET state histograms are on the right of each panel plotted as means of three independent experiments ± SEM. Assignments of the FRET states to the protein conformations are on the right of panel G. (hm) Dwell-time distributions of FRET states corresponding to: ‘full TM interface’ (high FRET for E385C, Swap or Turn) (H) or ‘reduced TM interface’ (low FRET for E385C, Twist) (I) in 100 mM Cl or 0 Cl (L and M, respectively); αA bound (high FRET for Q24C, Swap) (J) or αA loose (low FRET for Q24C, Turn and Twist) (K) in 0 mM Cl-. Dashed lines are the fluorophore survival probability, which refers to the fraction of molecules where both LD555 and D655 survived longer than the time indicated in the x-axis at pH 7.5 (blue) and pH 4.5 (red). Data are averages of three independent repeats, and error bars are SEM.

Source data

Extended Data Fig. 7 Examples of smFRET recordings of CLC-ec1.

(a-h) Representative smFRET traces of CLC-ec1 labelled at E385C at pH 7.5 and 0 mM Cl (A), pH 7.5 and 100 mM Cl (B), pH 4.5 and 0 mM Cl (C), pH 4.5 and 100 mM Cl (D), or CLC-ec1 labelled at Q24C at pH 7.5 and 0 mM Cl (E), pH 7.5 and 100 mM Cl (F), pH 4.5 and 0 mM Cl (G) and pH 4.5 and 100 mM Cl (H). Fluorescence emissions from LD555p (green) and LD655 (red) dyes are shown in the top panels, and FRET efficiency (blue) is in the bottom panels. The traces were selected to illustrate the heterogeneity of the dynamic behaviors observed in all collected datasets. The quantitative analysis of the FRET-efficiency and dwell time distributions is reported in Fig. 3 and Extended Data Fig. 6.

Source data

Extended Data Fig. 8 Water wire formation and Cl pathway dynamics of CLC-ec1.

(a) The pKa values of the ionizable residues calculated using PropKa3,4. Black dotted lines correspond to pH 4.5 and 7.5. Data are not shown for residues with the calculated pKa values below 4.5 or above 7.5 in all conformations. (b-f) Spontaneous formation of water wires connecting Gluex and the intracellular bulk water within the H+ pathways of protomers A and B, observed in 10 independent 1 μs long simulations of CLC-ec1 WT in Swap (B), Turn (C), Twist (D), Bot Intermediate crosslinked (E), and E202Y (F). Each vertical line in the panels shows the occurrence of a water wire at that time point. The number of events of formation of water wires as a function of the lifetime of individual event. The lifetime of water wire is defined as the time duration for any hydrogen bond in the water wire to be disconnected. (H)Representative pore-radius profiles along the z-axis of the Cl pathway, calculated using HOLE1 for conformations with the open (green) and closed (blue) external gate seen in the MD simulations of Twist. (i) The diameter of the Cl permeation pathway, visualized using HOLE1 in MD snapshots of Twist with an open (right panel) and closed (left panel) extracellular gate. The external gate radius is calculated using the backbone atoms of residues 147, 148, 356, and 357. Residues S107, E148, I356, and Y445 are shown as sticks. Blue indicates regions with R>2.3 Å, green 2.3 Å>R>1.15 Å, and red R< 1.15 Å. (j) Representative snapshots of conformational rearrangements in the backbone of the Cl pathway when the external gate is open (green) or closed (blue). * denotes the center of the Cl permeation pathway.

Source data

Extended Data Fig. 9 Characterization of activating mutations in CLC-ec1 and CLC-7.

(a) Theoretical (top panels - dimer & monomer) and experimental (middle and bottom panels – WT & ΔN29 pH 7.5 & 4.5) photobleaching steps probability distributions (P1, P2, P3+) for CLC-ec1 in lipid bilayers as a function of the experimental subunit/lipid mole fraction, χobs. Data for WT and ΔN29 pH 7.5 & 4.5 are means ± SEM for n = 4 for WT pH 4.5 and n = 3 for all other systems, where n is the number of independently prepared samples. (b, d) Representative normalized time courses of chloride efflux from liposomes reconstituted at pH 4.5 (B) or 7.5 (D) with CLC-ec1 WT (black), Q24C (orange), I201W (purple), E202Q (red) and E202Y (grey) at 0.2 μg protein/mg lipid. Traces were normalized to the total chloride content of the liposomes determined following the addition of detergent2. (c-e) Initial velocity of Cl efflux at pH 4.5 (C) and 7.5 (E) of proteoliposomes reconstituted with CLC-ec1 WT (white), Q24C (orange), I201W (purple), E202Q (red) and E202Y (grey). All values are shown as means ± S.D. The number of repeats of independent experiments and preparations of proteo-liposome samples for all constructs are reported in Supplementary Table 1. (f, g) smFRET population contour plots of CLC-ec1 E202Q (F) and E202Y (G) labelled at E385C at pH 7.5 (left) and 4.5 (right) in 100 mM Cl. Data are average of 3 independent repeats, and errors are SEM. (h) Forward and reverse potential of mean force (PMF) profile as a function of the minimum distance between αA of monomer A and αR of monomer B with E202 protonated, E202(0), or deprotonated, E202(-), at pH 4.5 or pH 7.5 with no Cl in the central binding site. Molecular systems are listed in Supplementary Table 4. Error bars (in SD) are estimated by calculating the PMF values in four consecutive blocks of the trajectories. (i) Free energy difference between local energy minima for αA ‘bound’ to and ‘loose’ from αR with E202 protonated (E2020) or deprotonated (E202) at pH 4.5 or pH 7.5 with no Cl in the central binding site. ΔG is taken from the free energy profiles in panel H. Error bars are estimated in the same fashion as in panel H. (j) Current density-voltage curves for CLC-7PM WT (black), E311Q (red), L310W (blue), G578W (pink), and L310W/G578W (WW, orange). Data are averages of 10 independent repeats, and error bars are SEM.

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1 and 2 and Supplementary Tables 1–7.

Reporting Summary

Peer Review File

Supplementary Data 1

Primer sequences used in mutagenesis

Supplementary Video 1

The pH and Cl dependent rearrangements of the CLC-ec1 dimer interface are viewed from the intracellular side. Morph between Swap, Turn, Twist conformations of CLC-ec1 viewed from the intracellular solution. The transmembrane dimer interface helices αI, αH, αP and αQ are shown in light pink, and the other helices of each protomer are colored by white and yellow, respectively.

Supplementary Video 2

The pH and Cl dependent rearrangements of the CLC-ec1 dimer are viewed from the plane of the membrane. The same morph as in Supplementary Video 1, but viewed from the plane of the membrane.

Source data

Source Data Fig. 1

Source data of all graphs in Fig. 1.

Source Data Fig. 2

Source data of all graphs in Fig. 2.

Source Data Fig. 3

Source data of all graphs in Fig. 3.

Source Data Fig. 4

Source data of all graphs in Fig. 4.

Source Data Fig. 5

Source data of all graphs in Fig. 5.

Source Data Extended Data Fig. 2

Source data used in all graphs in Extended Data Fig. 2A,B,C.

Source Data Extended Data Fig. 3

Source data used in all graphs in Extended Data Fig. 3A,B,F,G,H.

Source Data Extended Data Fig. 4

Source data used in all graphs in Extended Data Fig. 4C.

Source Data Extended Data Fig. 4

Uncropped gel image for Extended Data Fig. 4A,B.

Source Data Extended Data Fig. 5

Source data used in all graphs in Extended Data Fig. 5A,B,C.

Source Data Extended Data Fig. 6

Source data used in all graphs in Extended Data Fig. 6A,B,D,E,F,G,H,I,J,K,L,M.

Source Data Extended Data Fig. 7

Source data used in all graphs in Extended Data Fig. 7.

Source Data Extended Data Fig. 8

Source data used in all graphs in Extended Data Fig. 8A,B,G,H.

Source Data Extended Data Fig. 9

Source data used in all graphs in Extended Data Fig. 9A,B,C,D,E,F,G,H,I,J.

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

Fortea, E., Lee, S., Chadda, R. et al. Structural basis of pH-dependent activation in a CLC transporter. Nat Struct Mol Biol 31, 644–656 (2024). https://doi.org/10.1038/s41594-023-01210-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-023-01210-5

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