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:

Discrimination between cyclic nucleotides in a cyclic nucleotide-gated ion channel

Abstract

Cyclic nucleotide-gated ion channels are crucial in many physiological processes such as vision and pacemaking in the heart. SthK is a prokaryotic homolog with high sequence and structure similarities to hyperpolarization-activated and cyclic nucleotide-modulated and cyclic nucleotide-gated channels, especially at the level of the cyclic nucleotide binding domains (CNBDs). Functional measurements showed that cyclic adenosine monophosphate (cAMP) is a channel activator while cyclic guanosine monophosphate (cGMP) barely leads to pore opening. Here, using atomic force microscopy single-molecule force spectroscopy and force probe molecular dynamics simulations, we unravel quantitatively and at the atomic level how CNBDs discriminate between cyclic nucleotides. We find that cAMP binds to the SthK CNBD slightly stronger than cGMP and accesses a deep-bound state that a cGMP-bound CNBD cannot reach. We propose that the deep binding of cAMP is the discriminatory state that is essential for cAMP-dependent channel activation.

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: 2D crystallization of SthK CL-CNBD.
Fig. 2: Schematic of the experiment to probe the cN–CNBD interaction using AFM-SMFS.
Fig. 3: Binding kinetics of cAMP–CNBD and cGMP–CNBD.
Fig. 4: H-bond interactions between CNBD and cNs.
Fig. 5: A deep cAMP binding site after extended cAMP–CNBD bond formation.
Fig. 6: Kinetic model of cAMP and cGMP binding to the SthK CNBD.

Similar content being viewed by others

Data availability

The source data files contain all data (force distribution histograms, H-bond distributions) necessary to interpret, verify and extend the presented work. In the absence of dedicated data repositories for raw data AFM force curves and MDS trajectories, and in light of the instructions needed to open these files in proprietary software (in the case of the AFM force curves) and the additional information (parameters and conditions) needed to understand and use the data, raw data AFM force curves and MDS trajectories can be received from S.S. (sis2019@med.cornell.edu) and H.G. (hgrubmu@gwdg.de), respectively, upon reasonable request. Source data are provided with this paper.

References

  1. Kaupp, U. B. & Seifert, R. Cyclic nucleotide-gated ion channels. Physiol. Rev. 82, 769–824 (2002).

    Article  CAS  PubMed  Google Scholar 

  2. Craven, K. B. & Zagotta, W. N. CNG and HCN channels: two peas, one pod. Annu. Rev. Physiol. 68, 375–401 (2006).

    Article  CAS  PubMed  Google Scholar 

  3. Robinson, R. B. & Siegelbaum, S. A. Hyperpolarization-activated cation currents: from molecules to physiological function. Annu. Rev. Physiol. 65, 453–480 (2003).

    Article  CAS  PubMed  Google Scholar 

  4. Biel, M., Wahl-Schott, C., Michalakis, S. & Zong, X. Hyperpolarization-activated cation channels: from genes to function. Physiol. Rev. 89, 847–885 (2009).

    Article  CAS  PubMed  Google Scholar 

  5. James, Z. M. & Zagotta, W. N. Structural insights into the mechanisms of CNBD channel function. J. Gen. Physiol. 150, 225–244 (2017).

    Article  PubMed  Google Scholar 

  6. Yu, F. H., Yarov-Yarovoy, V., Gutman, G. A. & Catterall, W. A. Overview of molecular relationships in the voltage-gated ion channel superfamily. Pharmacol. Rev. 57, 387–395 (2005).

    Article  CAS  PubMed  Google Scholar 

  7. Rheinberger, J., Gao, X., Schmidpeter, P. A. M. & Nimigean, C. M. Ligand discrimination and gating in cyclic nucleotide-gated ion channels from apo and partial agonist-bound cryo-EM structures. eLife 7, e39775 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  8. Lee, C.-H. & MacKinnon, R. Structures of the human HCN1 hyperpolarization-activated channel. Cell 168, 111–120.e111 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Saponaro, A. et al. Gating movements and ion permeation in HCN4 pacemaker channels. Mol. Cell 81, 2929–2943.e2926 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Li, M. et al. Structure of a eukaryotic cyclic-nucleotide-gated channel. Nature 542, 60–65 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Schmidpeter, P. A. M., Gao, X., Uphadyay, V., Rheinberger, J. & Nimigean, C. M. Ligand binding and activation properties of the purified bacterial cyclic nucleotide–gated channel SthK. J. Gen. Physiol. 150, 821–834 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Marchesi, A. et al. An iris diaphragm mechanism to gate a cyclic nucleotide-gated ion channel. Nat. Commun. 9, 3978 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Kesters, D. et al. Structure of the SthK carboxy-terminal region reveals a gating mechanism for cyclic nucleotide-modulated Ion channels. PLoS ONE 10, e0116369 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Florin, E. L., Moy, V. T. & Gaub, H. E. Adhesion forces between individual ligand-receptor pairs. Science 264, 415 (1994).

    Article  CAS  PubMed  Google Scholar 

  15. Thess, A. et al. Specific orientation and two-dimensional crystallization of the proteasome at metal-chelating lipid interfaces. J. Biol. Chem. 277, 36321–36328 (2002).

    Article  CAS  PubMed  Google Scholar 

  16. Nye, J. A. & Groves, J. T. Kinetic control of histidine-tagged protein surface density on supported lipid bilayers. Langmuir 24, 4145–4149 (2008).

    Article  CAS  PubMed  Google Scholar 

  17. Lolicato, M. et al. Tetramerization dynamics of C-terminal domain underlies isoform-specific cAMP gating in hyperpolarization-activated cyclic nucleotide-gated channels. J. Biol. Chem. 286, 44811–44820 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Zagotta, W. N. et al. Structural basis for modulation and agonist specificity of HCN pacemaker channels. Nature 425, 200–205 (2003).

    Article  CAS  PubMed  Google Scholar 

  19. Lolicato, M. et al. Cyclic dinucleotides bind the C-linker of HCN4 to control channel cAMP responsiveness. Nat. Chem. Biol. 10, 457–462 (2014).

    Article  CAS  PubMed  Google Scholar 

  20. Ebner, A. et al. A new, simple method for linking of antibodies to atomic force microscopy tips. Bioconjug. Chem. 18, 1176–1184 (2007).

    Article  CAS  PubMed  Google Scholar 

  21. Baumgartner, W. et al. Cadherin interaction probed by atomic force microscopy. Proc. Natl Acad. Sci. USA 97, 4005–4010 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Evans, E. & Ritchie, K. Dynamic strength of molecular adhesion bonds. Biophys. J. 72, 1541–1555 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Huang, J. et al. The kinetics of two-dimensional TCR and pMHC interactions determine T-cell responsiveness. Nature 464, 932–936 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Lo Giudice, C., Zhang, H., Wu, B. & Alsteens, D. Mechanochemical activation of class-B G-protein-coupled receptor upon peptide–ligand binding. Nano Lett. 20, 5575–5582 (2020).

    Article  CAS  PubMed  Google Scholar 

  25. Zhu, R. et al. Allosterically linked binding sites in serotonin transporter revealed by single molecule force spectroscopy. Front. Mol. Biosci. 7, 99 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Williams, P. M. Analytical descriptions of dynamic force spectroscopy: behaviour of multiple connections. Anal. Chim. Acta 479, 107–115 (2003).

    Article  CAS  Google Scholar 

  27. Rankl, C. et al. Multiple receptors involved in human rhinovirus attachment to live cells. Proc. Natl Acad. Sci. USA 105, 17778–17783 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Sulchek, T. A. et al. Dynamic force spectroscopy of parallel individual Mucin1–antibody bonds. Proc. Natl Acad. Sci. USA 102, 16638–16643 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. White, D. S. et al. cAMP binding to closed pacemaker ion channels is non-cooperative. Nature 595, 606–610 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Schmidpeter, P. A. M., Rheinberger, J. & Nimigean, C. M. Prolyl isomerization controls activation kinetics of a cyclic nucleotide-gated ion channel. Nat. Commun. 11, 6401 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Dudko, O. K., Hummer, G. & Szabo, A. Intrinsic rates and activation free energies from single-molecule pulling experiments. Phys. Rev. Lett. 96, 108101 (2006).

    Article  PubMed  Google Scholar 

  32. Aman, T. K., Gordon, S. E. & Zagotta, W. N. Regulation of CNGA1 channel gating by interactions with the membrane. J. Biol. Chem. 291, 9939–9947 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Brady, J. D. et al. Interplay between PIP3 and calmodulin regulation of olfactory cyclic nucleotide-gated channels. Proc. Natl Acad. Sci. USA 103, 15635–15640 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Asakawa, H., Yoshioka, S., Nishimura, K.-I. & Fukuma, T. Spatial distribution of lipid headgroups and water molecules at membrane/water interfaces visualized by three-dimensional scanning force microscopy. ACS Nano 6, 9013–9020 (2012).

  35. Koehler, M. et al. Control of ligand-binding specificity using photocleavable linkers in AFM force spectroscopy. Nano Lett. 20, 4038–4042 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Miyagi, A. & Scheuring, S. Automated force controller for amplitude modulation atomic force microscopy. Rev. Sci. Instrum. 87, 053705 (2016).

    Article  PubMed  Google Scholar 

  37. Miyagi, A. & Scheuring, S. A novel phase-shift-based amplitude detector for a high-speed atomic force microscope. Rev. Sci. Instrum. 89, 083704 (2018).

    Article  PubMed  Google Scholar 

  38. Sali, A. & Blundell, T. L. Comparative protein modelling by satisfaction of spatial restraints. J. Mol. Biol. 234, 779–815 (1993).

    Article  CAS  PubMed  Google Scholar 

  39. Wang, J., Wolf, R. M., Caldwell, J. W., Kollman, P. A. & Case, D. A. Development and testing of a general amber force field. J. Comput. Chem. 25, 1157–1174 (2004).

    Article  CAS  PubMed  Google Scholar 

  40. Wang, J., Wang, W., Kollman, P. A. & Case, D. A. Automatic atom type and bond type perception in molecular mechanical calculations. J. Mol. Graph Model 25, 247–260 (2006).

    Article  PubMed  Google Scholar 

  41. Bayly, C. I. et al. A well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP model. J. Phys. Chem. 97, 10269–10280 (1993).

    Article  CAS  Google Scholar 

  42. Frisch, M. J. et al. Gaussian v.3, Revision C.02 (Gaussian, Inc., 2004).

  43. Roothaan, C. C. J. New developments in molecular orbital theory. Rev. Mod. Phys. 23, 69–89 (1951).

    Article  CAS  Google Scholar 

  44. Rassolov, V. A., Pople, J. A., Ratner, M. A. & Windus, T. L. 6-31G* basis set for atoms K through Zn. J. Chem. Phys. 109, 1223–1229 (1998).

    Article  CAS  Google Scholar 

  45. Mobley, D. L., Chodera, J. D. & Dill, K. A. On the use of orientational restraints and symmetry corrections in alchemical free energy calculations. J. Chem. Phys. 125, 084902–084902 (2006).

    Article  PubMed  Google Scholar 

  46. Sousa da Silva, A. W. & Vranken, W. F. ACPYPE—AnteChamber PYthon Parser interfacE. BMC Res. Notes 5, 367 (2012).

  47. Larsson, P., Kneiszl, R. C. & Marklund, E. G. MkVsites: a tool for creating GROMACS virtual sites parameters to increase performance in all-atom molecular dynamics simulations. J. Comput. Chem. 41, 1564–1569 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. 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).

    Article  CAS  Google Scholar 

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

    Article  Google Scholar 

  50. Verlet, L. Computer ‘experiments’ on classical fluids. I. Thermodynamical properties of Lennard-Jones molecules. Phys. Rev. 159, 98–103 (1967).

    Article  CAS  Google Scholar 

  51. Lindorff-Larsen, K. et al. Improved side-chain torsion potentials for the Amber ff99SB protein force field. Proteins 78, 1950–1958 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. 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).

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

    Article  CAS  Google Scholar 

  54. Vriend, G. WHAT IF: a molecular modeling and drug design program. J. Mol. Graph 8, 52–56 (1990).

    Article  CAS  PubMed  Google Scholar 

  55. Bussi, G., Donadio, D. & Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 126, 014101 (2007).

    Article  PubMed  Google Scholar 

  56. Berendsen, H. J. C., Postma, J. P. M., Gunsteren, W. F. V., DiNola, A. & Haak, J. R. Molecular dynamics with coupling to an external bath. J. Chem. Phys. 81, 3684–3690 (1984).

    Article  CAS  Google Scholar 

  57. Espinosa, E. & Molins, E. Retrieving interaction potentials from the topology of the electron density distribution: the case of hydrogen bonds. J. Chem. Phys. 113, 5686–5694 (2000).

    Article  CAS  Google Scholar 

  58. Evans, E. Probing the relation between force—lifetime—and chemistry in single molecular bonds. Annu. Rev. Biophys. Biomol. Struct. 30, 105–128 (2001).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank M. Rangl for initial experiments. Work in the Scheuring laboratory was supported by grants from the National Institute of Health, National Center for Complementary and Integrative Health DP1AT010874 and National Institute of Neurological Disorders and Stroke R01NS110790, and by the Kavli Institute at Cornell. Work in the Nimigean laboratory was supported by the National Institute of Health (GM124451 to C.M.N.) and the American Heart Association (18POST33960309 to P.A.M.S.). Work in the Grubmüller laboratory was funded by the Max Planck Society and by the German Science Foundation (DFG), Excellence Strategy Grant MBExC 2067/1-390729940; computer time was provided by the Max Planck Computing and Data Facility. A.C.V. was additionally supported by the European Union’s Horizon 2020 Framework Programme for Research and Innovation, Specific Grant Agreement No. 945539 (Human Brain Project SGA3).

Author information

Authors and Affiliations

Authors

Contributions

Y.P., C.M.N. and S.S. designed the experiments. P.A.M.S. expressed and purified the protein. Y.P. performed HS-AFM and AFM-SMFS experiments. Y.P. did the HS-AFM and AFM-SMFS data analysis. E.P., A.C.V. and H.G. performed and analyzed force probe MDS. Y.P., E.P., H.G. and S.S. wrote the paper. All authors edited the paper.

Corresponding author

Correspondence to Simon Scheuring.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Structural & Molecular Biology thanks Baron Chanda and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available. Primary Handling Editors: Florian Ullrich and Katarzyna Ciazynska, 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 cNs-CNBD binding studies using microscale thermophoresis (MST).

a) and c) Microscale thermophoresis (MST) traces of cAMP-CNBD (a) and cGMP-CNBD (c) interactions, respectively. b) Binding curve of labeled CNBD with cAMP. Two binding phases were detected. The first had a KD of 0.4 ± 0.5 µM, and the second had a KD of 1.6 ± 1.1 µM. d) Binding curve of labeled CNBD with cGMP. The KD value of cGMP-CNBD was 3.3 ± 1.8 µM. The concentration of labeled His6-C-linker-CNBD was kept constant (50 nM), and the concentration of cNs was varied from 0.00305 µM to 100 µM.

Source data

Extended Data Fig. 2 Distributions of rupture forces of cAMP-CNBD or cGMP-CNBD bonds under various conditions.

a) Force distribution of cAMP-CNBD unbinding following 0.02 s bond formation at varying pulling velocity (top right in each graph). b) Force distribution of cAMP-CNBD unbinding following 1.00 s bond formation at varying pulling velocities (top right in each graph). The force distributions are fitted with bimodal Gaussian fits to extract the most probable rupture forces for binding state 1 (first peak) and binding state 2 (second peak), which are then used for Bell–Evans model fitting. c) Force distribution of cGMP-CNBD unbinding following 0.02 s bond formation at varying pulling velocities (top right in each graph). d) Force distribution of cGMP-CNBD unbinding following 1.00 s bond formation at varying pulling velocities (top right in each graph). e) and f) Distributions of rupture forces of cGMP-CNBD (e) and cAMP-CNBD (f) unbinding following different bond formation times and at 0.4 µm/s pulling velocity). Bimodal Gaussian fits are used to extract the number of events for each bond state, which is then used to calculate the probability of occurrence of binding state 2.

Source data

Extended Data Fig. 3 Deviations of rupture force distributions.

Theoretical, simulation and experimental unbinding force distributions (normalized by loading rates) of cAMP (left) and cGMP (right). The experimental unbinding force distributions are the distributions after 0.02 s bond formation time.

Source data

Extended Data Fig. 4 Number of bond ruptures as a function of bond-formation time for cAMP-CNBD (left) and cGMP-CNBD (right).

With increasing cN contact time, the frequency of unsuccessful (0 bond) force-distance cycles decreases, while the number of successful (1 bond) force-distance cycles increases. Concomitantly, a fraction of force-distance cycles reported multiple (2 or 3) binding events.

Source data

Extended Data Fig. 5 Markovian sequence analysis for the force spectroscopy experiments of cAMP-CNBD at 1 s contact time.

Markovian model fitting for 1 (gray line) and 2 (dashed line) bonds. koff and xβ values were derived from the Bell–Evans model fit to the canonical binding mode (see Fig. 5b). The most probable rupture force (Gaussian peak) and error (full width at half maximum of the Gaussian peak) at each loading rate was determined through Gaussian fitting of the corresponding histogram (total data points for all histograms, N = 1898, Extended Data Fig. 2b).

Extended Data Fig. 6 Comparison of SthK with CNG and HCN channels, with a focus on their cN binding pockets.

Top: cNs binding pocket of SthK (left, gray) and CNGA1 (right, orange) showing that the residues Y and F are swapped in CNGA1 channels compared with SthK. Bottom: Sequence alignment of SthK with CNG and HCN channels showing that Y357 and F365 identified by MDS to be crucial in cN discrimination in SthK are not conserved, whereas M369 is a conservative mutation and A370 is a semi-conservative mutation. cAMP interacts with Y357, M369 and A370, and cGMP interacts with F365.

Supplementary information

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. 4

Statistical source data.

Source Data Fig. 5

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.

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

Pan, Y., Pohjolainen, E., Schmidpeter, P.A.M. et al. Discrimination between cyclic nucleotides in a cyclic nucleotide-gated ion channel. Nat Struct Mol Biol 30, 512–520 (2023). https://doi.org/10.1038/s41594-023-00955-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-023-00955-3

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