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:

Evolutionary balance between foldability and functionality of a glucose transporter

Abstract

Despite advances in resolving the structures of multi-pass membrane proteins, little is known about the native folding pathways of these complex structures. Using single-molecule magnetic tweezers, we here report a folding pathway of purified human glucose transporter 3 (GLUT3) reconstituted within synthetic lipid bilayers. The N-terminal major facilitator superfamily (MFS) fold strictly forms first, serving as a structural template for its C-terminal counterpart. We found polar residues comprising the conduit for glucose molecules present major folding challenges. The endoplasmic reticulum membrane protein complex facilitates insertion of these hydrophilic transmembrane helices, thrusting GLUT3’s microstate sampling toward folded structures. Final assembly between the N- and C-terminal MFS folds depends on specific lipids that ease desolvation of the lipid shells surrounding the domain interfaces. Sequence analysis suggests that this asymmetric folding propensity across the N- and C-terminal MFS folds prevails for metazoan sugar porters, revealing evolutionary conflicts between foldability and functionality faced by many multi-pass membrane proteins.

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: Single-molecule magnetic tweezers assay for observing folding of single GLUT3 proteins.
Fig. 2: Identification of the folding order of GLUT3 domains.
Fig. 3: Dissection of the folding order of individual MFS folds.
Fig. 4: EMC facilitates folding of single GLUT3 proteins.
Fig. 5: PE lipids facilitate GLUT3 inter-domain assembly.
Fig. 6: Phylogenetic analysis reveals an asymmetric folding propensity across the N- and C-domains common for metazoan sugar transporters.

Similar content being viewed by others

Data availability

All data that support the findings of this study are available in the manuscript or supplementary figures. Raw data have been deposited in Github (https://github.com/tyyoonlab/ Nat_Chem_biol_NCHEMB - A210813554). The following PDB IDs were used (4ZWB, 4ZWC, 4ZW9, 4GBZ, 6RW3, 4LDS and 6H7D). The following UniProt IDs were used (P11169, P0AGF4, O97467, A0A0H2VG78 and Q9LT15). Source data are provided with this paper.

Code availability

A program, written in LabView, to control the magnetic tweezers apparatus has been deposited in Github (https://github.com/tyyoonlab/Science_aaw8208). Codes for analyzing the magnetic tweezers and FACS data has been deposited in Github (https://github.com/tyyoonlab/Nat_Chem_biol_NCHEMB-A210813554). Codes for estimating the helix insertion energy are available at Github under GPL3.0 license (https://github.com/schnamo/TMH_insertion_energy).

References

  1. Hediger, M. A. et al. The ABCs of solute carriers: physiological, pathological and therapeutic implications of human membrane transport proteins. Pflügers Arch. 447, 465–468 (2004).

    Article  CAS  PubMed  Google Scholar 

  2. Cheng, Y. Membrane protein structural biology in the era of single particle cryo-EM. Curr. Opin. Struct. Biol. 52, 58–63 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Marinko, J. T. et al. Folding and misfolding of human membrane proteins in health and disease: from single molecules to cellular proteostasis. Chem. Rev. 119, 5537–5606 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Guna, A. & Hegde, R. S. Transmembrane domain recognition during membrane protein biogenesis and quality control. Curr. Biol. 28, R498–R511 (2018).

    Article  CAS  PubMed  Google Scholar 

  5. Oberai, A., Ihm, Y., Kim, S. & Bowie, J. U. A limited universe of membrane protein families and folds. Protein Sci. 15, 1723–1734 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Shi, Y. Common folds and transport mechanisms of secondary active transporters. Annu. Rev. Biophys. 42, 51–72 (2013).

    Article  PubMed  Google Scholar 

  7. Chitwood, P. J. & Hegde, R. S. An intramembrane chaperone complex facilitates membrane protein biogenesis. Nature 584, 630–634 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Kota, J. & Ljungdahl, P. O. Specialized membrane-localized chaperones prevent aggregation of polytopic proteins in the ER. J. Cell Biol. 168, 79–88 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Choi, H.-K. et al. Watching helical membrane proteins fold reveals a common N-to-C-terminal folding pathway. Science 366, 1150–1156 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Pao, S. S., Paulsen, I. T. & Saier, M. H. Jr. Major facilitator superfamily. Microbiol. Mol. Biol. Rev. 62, 1–34 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Yan, N. Structural biology of the major facilitator superfamily transporters. Annu. Rev. Biophys. 44, 257–283 (2015).

    Article  CAS  PubMed  Google Scholar 

  12. Deng, D. et al. Molecular basis of ligand recognition and transport by glucose transporters. Nature 526, 391–396 (2015).

    Article  CAS  PubMed  Google Scholar 

  13. Madej, M. G. Function, structure, and evolution of the major facilitator superfamily: the LacY manifesto. Adv. Biol. 2014, 1–20 (2014).

    Article  Google Scholar 

  14. Drew, D. & Boudker, O. Shared molecular mechanisms of membrane transporters. Annu. Rev. Biochem. 85, 543–572 (2016).

    Article  CAS  PubMed  Google Scholar 

  15. Hingorani, K. S. & Gierasch, L. M. Comparing protein folding in vitro and in vivo: foldability meets the fitness challenge. Curr. Opin. Struct. Biol. 24, 81–90 (2014).

    Article  CAS  PubMed  Google Scholar 

  16. Min, D., Arbing, M. A., Jefferson, R. E. & Bowie, J. U. A simple DNA handle attachment method for single molecule mechanical manipulation experiments. Protein Sci. 25, 1535–1544 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Findlay, H. E. & Booth, P. J. The biological significance of lipid–protein interactions. J. Phys. Condens. Matter 18, S1281–S1291 (2006).

    Article  CAS  PubMed  Google Scholar 

  18. Ujwal, R. & Bowie, J. U. Crystallizing membrane proteins using lipidic bicelles. Methods 55, 337–341 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Findlay, H. E. & Booth, P. J. The folding, stability and function of lactose permease differ in their dependence on bilayer lipid composition. Sci. Rep. 7, 13056 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Hessa, T. et al. Molecular code for transmembrane-helix recognition by the Sec61 translocon. Nature 450, 1026–1030 (2007).

    Article  CAS  PubMed  Google Scholar 

  21. Snider, C., Jayasinghe, S., Hristova, K. & White, S. H. MPEx: a tool for exploring membrane proteins. Protein Sci. 18, 2624–2628 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Moon, C. P. & Fleming, K. G. Side-chain hydrophobicity scale derived from transmembrane protein folding into lipid bilayers. Proc. Natl Acad. Sci. USA 108, 10174–10177 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Lee, T.-H. Extracting kinetics information from single-molecule fluorescence resonance energy transfer data using hidden Markov models. J. Phys. Chem. B 113, 11535–11542 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Zhang, Y., Jiao, J. & Rebane, A. A. Hidden Markov modeling with detailed balance and its application to single protein folding. Biophys. J. 111, 2110–2124 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Radestock, S. & Forrest, L. R. The alternating-access mechanism of MFS transporters arises from inverted-topology repeats. J. Mol. Biol. 407, 698–715 (2011).

    Article  CAS  PubMed  Google Scholar 

  26. Jonikas, M. C. et al. Comprehensive characterization of genes required for protein folding in the endoplasmic reticulum. Science 323, 1693–1697 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Pleiner, T. et al. Structural basis for membrane insertion by the human ER membrane protein complex. Science 369, 433–436 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Chitwood, P. J., Juszkiewicz, S., Guna, A., Shao, S. & Hegde, R. S. EMC is required to initiate accurate membrane protein topogenesis. Cell 175, 1507–1519.e16 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Guna, A., Volkmar, N., Christianson, J. C. & Hegde, R. S. The ER membrane protein complex is a transmembrane domain insertase. Science 359, 470–473 (2018).

    Article  CAS  PubMed  Google Scholar 

  30. Kauko, A. et al. Repositioning of transmembrane α-helices during membrane protein folding. J. Mol. Biol. 397, 190–201 (2010).

    Article  CAS  PubMed  Google Scholar 

  31. Lai, G. & Renthal, R. Integral membrane protein fragment recombination after transfer from nanolipoprotein particles to bicelles. Biochemistry 52, 9405–9412 (2013).

    Article  CAS  PubMed  Google Scholar 

  32. Jo, S., Lim, J. B., Klauda, J. B. & Im, W. CHARMM-GUI Membrane Builder for mixed bilayers and its application to yeast membranes. Biophys. J. 97, 50–58 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Cheung, M. S., García, A. E. & Onuchic, J. N. Protein folding mediated by solvation: water expulsion and formation of the hydrophobic core occur after the structural collapse. Proc. Natl Acad. Sci. USA 99, 685–690 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Bowie, J. U. Solving the membrane protein folding problem. Nature 438, 581–589 (2005).

    Article  CAS  PubMed  Google Scholar 

  35. Bogdanov, M. & Dowhan, W. Lipid-dependent generation of dual topology for a membrane protein. J. Biol. Chem. 287, 37939–37948 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Steinegger, M. et al. HH-suite3 for fast remote homology detection and deep protein annotation. BMC Bioinform. 20, 473 (2019).

    Article  Google Scholar 

  37. Seatter, M. J., De La Rue, S. A., Porter, L. M. & Gould, G. W. QLS motif in transmembrane helix VII of the glucose transporter family interacts with the C-1 position of D-glucose and is involved in substrate selection at the exofacial binding site. Biochemistry 37, 1322–1326 (1998).

    Article  CAS  PubMed  Google Scholar 

  38. Forrest, L. R. Structural symmetry in membrane proteins. Annu. Rev. Biophys. 44, 311–337 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Cymer, F. & von Heijne, G. Cotranslational folding of membrane proteins probed by arrest-peptide-mediated force measurements. Proc. Natl Acad. Sci. USA 110, 14640–14645 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Lolkema, J. S., Dobrowolski, A. & Slotboom, D.-J. Evolution of antiparallel two-domain membrane proteins: tracing multiple gene duplication events in the DUF606 family. J. Mol. Biol. 378, 596–606 (2008).

    Article  CAS  PubMed  Google Scholar 

  41. Walmsley, A. R., Barrett, M. P., Bringaud, F. & Gould, G. W. Sugar transporters from bacteria, parasites and mammals: structure–activity relationships. Trends Biochem. Sci. 23, 476–481 (1998).

    Article  CAS  PubMed  Google Scholar 

  42. Mueckler, M. & Thorens, B. The SLC2 (GLUT) family of membrane transporters. Mol. Asp. Med. 34, 121–138 (2013).

    Article  CAS  Google Scholar 

  43. Thorens, B. GLUT2, glucose sensing and glucose homeostasis. Diabetologia 58, 221–232 (2015).

    Article  CAS  PubMed  Google Scholar 

  44. Serdiuk, T. et al. YidC assists the stepwise and stochastic folding of membrane proteins. Nat. Chem. Biol. 12, 911–917 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Wang, F., Chan, C., Weir, N. R. & Denic, V. The Get1/2 transmembrane complex is an endoplasmic-reticulum membrane protein insertase. Nature 512, 441–444 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. McGilvray, P. T. et al. An ER translocon for multi-pass membrane protein biogenesis. eLife 9, e56889 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Hedge, R. S. & Keenan, R. J. The mechanisms of integral membrane protein biogenesis. Nat. Rev. Mol. Cell Biol. 23, 107–124 (2022).

    Article  Google Scholar 

  48. Zhang, W., Campbell, H. A., King, S. C. & Dowhan, W. Phospholipids as determinants of membrane protein topology: phosphatidylethanolamine is required for the proper topological organization of the γ-aminobutyric acid permease (GabP) of Escherichia coli. J. Biol. Chem. 280, 26032–26038 (2005).

    Article  CAS  PubMed  Google Scholar 

  49. Volmer, R. & Ron, D. Lipid-dependent regulation of the unfolded protein response. Curr. Opin. Cell Biol. 33, 67–73 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Jacquemyn, J., Cascalho, A. & Goodchild, R. E. The ins and outs of endoplasmic reticulum‐controlled lipid biosynthesis. EMBO Rep. 18, 1905–1921 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. O’Donnell, J. P. et al. The architecture of EMC reveals a path for membrane protein insertion. eLife 9, e57887 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Shon, M. J., Rah, S.-H. & Yoon, T.-Y. Submicrometer elasticity of double-stranded DNA revealed by precision force-extension measurements with magnetic tweezers. Sci. Adv. 5, eaav1697 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  53. Bouchiat, C. et al. Estimating the persistence length of a worm-like chain molecule from force-extension measurements. Biophys. J. 76, 409–413 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Oesterhelt, F. et al. Unfolding pathways of individual bacteriorhodopsins. Science 288, 143–146 (2000).

    Article  CAS  PubMed  Google Scholar 

  55. Seol, Y., Li, J., Nelson, P. C., Perkins, T. T. & Betterton, M. Elasticity of short DNA molecules: theory and experiment for contour lengths of 0.6–7 μm. Biophys. J. 93, 4360–4373 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Sarkar, A., Caamano, S. & Fernandez, J. M. The mechanical fingerprint of a parallel polyprotein dimer. Biophys. J. 92, L36–L38 (2007).

    Article  CAS  PubMed  Google Scholar 

  57. Gebhardt, J. C. M., Bornschlögl, T. & Rief, M. Full distance-resolved folding energy landscape of one single protein molecule. Proc. Natl Acad. Sci. USA 107, 2013–2018 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Hinczewski, M., von Hansen, Y. & Netz, R. R. Deconvolution of dynamic mechanical networks. Proc. Natl Acad. Sci. USA 107, 21493–21498 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Hinczewski, M., Gebhardt, J. C. M., Rief, M. & Thirumalai, D. From mechanical folding trajectories to intrinsic energy landscapes of biopolymers. Proc. Natl Acad. Sci. USA 110, 4500–4505 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Saier, M. H. Jr, Tran, C. V. & Barabote, R. D. TCDB: the transporter classification database for membrane transport protein analyses and information. Nucleic Acids Res. 34, D181–D186 (2006).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank E. Kweon for help with preparing illustrations. This work was supported by National Creative Research Initiative Program (NRF-2021R1A3B1071354 to T.-Y.Y.), the Bio Medical Technology Development Program (NRF-2018M3A9E2023523 to T.-Y.Y.) and NRF grants (NRF-2019M3E5D6063903 and NRF-2020R1A2C2003783 to H.-J.C.; NRF-2019R1A6A1A10073437 and NRF-2020M3A9G7103933 to M.S.; NRF-2021M3A9I4021220 and NRF-2020R1A6C101A183 to S.-H.R), all funded by the National Research Foundation of South Korea. This work was also supported by the UK Medical Research Council (MRC_UP_12-1/10 to E.A.M.), the US National Science Foundation (MCB-181069 to W.I.) and National Institutes of Health grant (R01GM118684 to H.H.).

Author information

Authors and Affiliations

Authors

Contributions

H.-K.C. and T.-Y.Y. conceived the project. H.-K.C., H.-J.C., E.A.M. and T.-Y.Y. designed the experiments. H.-K.C., C.L. and S.A.K. performed the magnetic tweezers experiment. H.K. and H.-J.C. expressed and purified GLUT3 proteins. B.P.P. and E.A.M. expressed and purified EMC. H.G.K. performed the flow cytometry experiment. S.P. and W.I. performed the molecular dynamics simulation. H.G.K., C.T. and M.S. performed the bioinformatic analysis. H.L. and S.-H.R. performed the TEM imaging. H.-K.C., C.L. and T.-Y.Y. prepared the manuscript, with assistance from H.-J.C., E.A.M. and H.H., and with input from all authors.

Corresponding authors

Correspondence to Elizabeth A. Miller, Hee-Jung Choi or Tae-Young Yoon.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Argyris Politis, Horst Vogel and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Sample preparation of WT and C-domain-knotted GLUT3 proteins and bicelle membranes.

a, Elution profile obtained by size exclusion chromatography (SEC) of WT GLUT3 (black) and S265C/A469C GLUT3 (red). b, Purified WT GLUT3 protein analyzed by SDS-PAGE. The major peak position in (a) was used for the gels. The right lane is molecular weight standards. c, Representative gel image of SDS-PAGE after SYBR green staining. The left lane shows SpyCatcher-DNA handle only, while the right lane exhibits a mixture of the SpyCatcher-DNA handle and the purified Spytag-GLUT3-spytag. d, Bicelle size as measured by dynamic light scattering (DLS) under indicated lipid compositions. The data with gray color is a control sample of 1-µm polystyrene beads. e, Mean diameter of the bicelles determined for each lipid composition. Error bars represent SEM (n = 22 and 25 for 30 mol% PG and 100 mol% PG, respectively). f,g, Electron microscope images of bicelles with different lipid compositions. Two specified electron microscopy methods are used. h, Representative trace for the force-jump experiment to determine the refolding probability. Force was first increased to 25pN to induce full unraveling of the protein and then relaxed to 1 pN for 500 s before checking the folding status through re-application of 25 pN. The folding status was determined by the unfolding steps observed under 25pN. i, Probability for observing the completely folded state for indicated buffer conditions. The refolding probability virtually abolished when GLUT3 was embedded in DDM micelles at 0.1 % (w/v), indicating that the lipid bilayer environments provided by the bicelle membranes are essential for inducing the fully folded state.

Source data

Extended Data Fig. 2 Analysis of single-molecule magnetic tweezer data.

a, Precision in determination of the vertical position of a bead as a function of the measurement bandwidth. The plot indicates an ~1 nm resolution when bead positions are averaged over 50 ms (~20 Hz sampling). In our magnetic tweezer experiments, the bicelle phase used for providing the lipid bilayer environments to the target membrane proteins offers additional low-frequency fluctuations, forcing a longer averaging time of 200 ms to achieve the 1 nm accuracy in our membrane protein folding studies. b, Bayesian Information Criteria (BIC) values of WT GLUT3 for each number of states with different bicelle compositions (n = 16 and 11 for 30 mol% and 100 mol% PG, respectively). c, Position of folding intermediates determined by HMM with for different number of states assumed in HMM analysis. The positions of the key intermediates (Uh, If4 (N-domain folded), If6 (C-domain folded) and N) are essentially preserved when the number of assumed states are changed, which only generats additional intermediates either in the middle of either N-domain folding or domain-domain assembly (blue arrows). Thus, our HMM analysis does not randomly assign the intermediate positions out of noisy data, but rather identifies the intermediates implicated in our time-resolved traces in a robust way. d, e, Representative folding traces for WT GLUT3 with 30 mol% PG (d) and 100 mol% PG (e) at 5pN. Two replicates are shown for each condition, and the gray and black traces are 1.2-kHz raw data and 5-Hz median-filtered data, respectively. Red traces indicate the transitions between intermediates identified by HMM.

Extended Data Fig. 3 Sample preparation and folding behavior of S265C/A469C GLUT3.

a, Schematic of the assay using BODIPY-L-cystine. The left panel is the chemical structure of two BODIPY FL fluorophores attached to the amino groups of the disulfide-containing amino acid, cysteine. The right panel shows the structure of GLUT3 before and after the treatment (addition of TCEP or increasing the temperature). Green dots in the right panel are the BODIPY FL fluorophores reacted with cysteines in GLUT3. b, Gel analysis for WT and S265C/A469C GLUT3 in the presence of TCEP. Upper gel shows the amount of GLUT3 stained by Coomassie blue. Lower gel shows the amount of BODIPY FL fluorophores reacted with cysteines in GLUT3. The stained positions are same in both gels. c, Fluorescence profile of BODIPY FL fluorophore-labeled GLUT3 as temperature increased. Dashed lines indicate the melting temperatures of the WT (black) and S265C/A469C GLUT3 (red). Error bars represent SEM (n = 4). d, A mechanical cycle for inducing refolding of a single S265C/A469C GLUT3(GLUT3CC) and corresponding structural states of the protein. e,f, Representative folding traces of GLUT3CC with 30 mol% PG (f) and 100 mol% PG (g) at 5 pN. Definitions of the traces are identical to those of Extended Data Fig. 2d. g, BIC values of GLUT3CC for each number of states (n = 22 and 12 for 30 mol% and 100 mol% PG, respectively).

Source data

Extended Data Fig. 4 Sample preparation and unfolding characteristics of T45C/K115C GLUT3.

a, Atomic contacts among TMHs 1, 2, and 4. Inset shows detailed position of interacting residues (blue for amino group, orange for carboxyl group, and yellow for thiol group). b, The positions of two mutations, T45C/K115C in GLUT3 (GLUT3TM23C). c, An absorbance profile of BODIPY FL fluorophore-labeled GLUT3TM23C as temperature increases. The experiment was done as depicted in Extended Data Fig. 3c. Error bars represent SEM (n = 4). d, Collection of 50Hz-median filtered unfolding traces initiated from N state for GLUT3TM23C. e, Distributions of extension values recorded during high-force unfolding of single GLUT3TM23C proteins. Extension values represent mean ± SD (n = 19).

Extended Data Fig. 5 Determination of folding order for N-domain of GLUT3.

a, Schematic of pulling geometry for N-domain of GLUT3 at 25 pN. dN-domain is the distance between two points of force application before unfolding (PDB: 4ZWC). ∆zi indicates the expected extension increase for GLUT3 for the ith intermediate. zi,p is the extension of the unfolded portion along the membrane for the ith intermediate. di denotes the distance between the points of force application for the ith intermediate. ∆ni is the number of amino acids of the unfolded portion. l is the length of a single amino acid. b, Unfolding extension distribution for the N-domain part of the WT GLUT3 at 25pN. c, Structural information and folding/unfolding order of the N-domain of WT GLUT3. The distance between two orange dashed lines (perpendicular to the membrane) represents the vertical distance between the two points of application (di). This orange dashed line forms an angle of \(\overline {\uptheta}\) with a black dashed line to the unfolded portion of N-domain in the membrane. d, Unfolding extension distribution for the N-domain part of GLUT3TM23C at 25pN. e, Structural information and folding/unfolding order of the N-domain of GLUT3TM23C. The description is the same as (c) except for protein construct. The disulfide bond of GLUT3TM23C did not affect the first unfolding step for N-domain that amounted to ~15.7 nm, confirming that TMHs 5 and 6 constitute the first unfolding step of N-domain. The second unfolding step was slightly reduced to 7.5 nm, which was consistent with the length of last helical turn of the long linker region when TMH 1 would be protected by knotting, mapping the second unfolding step to that of TMH 1 and its linker region. The last two unfolding steps before Uc were reduced to a single step of 4.2 nm, which would reflect unfolding of TMH 4 outside the knotted region. f, Representative traces showing the final unfolding step of a 4.2 nm extension increase for GLUT3TM23C. Three replicates are shown, and the value indicates the distance between two gaussian peaks.

Source data

Extended Data Fig. 6 Symmetrical structure of GLUT3.

a, Structural view of GLUT3’s N-domain with its C2 pseudo-symmetry. b, Structural view and electrostatic potential of the helix triplets composed of TMHs 1, 2, 3 and 4, 5, 6 each. c, Structural view and electrostatic potential of the helix triplets composed of TMHs 1, 5, 6 and 4, 2, 3 each.

Extended Data Fig. 7 EMC preparation and folding characteristics of GLUT3 with β2AR.

a, Structural model of the human ER membrane protein complex (EMC). b, Purified EMC analyzed by SDS-PAGE. The left lane is molecular weight standards and the right lane shows purified EMC. c, Fluorescence profile of EMC before and after freeze-thaw as the temperature is increased. Inset displays the profile for first derivative of fluorescence intensity. d, Representative folding traces of WT (left) and GLUT3CC (right) in the presence of EMC. The definition of each trace is identical to the traces in Extended Data Fig. 2d. e, Purified β2AR analyzed by SDS-PAGE. The left lane is molecular weight standards and the right lane shows β2AR. f, Representative folding traces for single GLUT3 at 5 pN with 30 mol% PG in the bicelles in the presence of 500 nM β2AR. Two replicates are shown. g, Probability distributions of deconvoluted extension values observed under indicated folding conditions at 5pN (n = 11 for the reaction with β2AR). The black and red distributions are revisited from Fig. 4e. The shaded area means SEM. h, BIC values for the indicated number of states (n = 13 and 11 traces for WT GLUT3 and S265C/A469C GLUT3, respectively). i, Positions of folding/unfolding intermediates identified with HMM are depicted for the indicated conditions. Error bars represent SEM (n = 22 and 35 traces for 100 mol% DMPG and 30 mol% DMPG with 500nM EMC, respectively).

Extended Data Fig. 8 Analysis of MD simulation for GLUT3 in lipid bilayer.

a, The average number of contacting water molecule to polar/charged residues in TMHs of N-domain with or without DMPE. Error bars represent SD (n = 4000 for each case). Polar/charged residues in TMHs of GLUT3 for the analysis are as follows. S21, Q23, N27, T28, S71, S78, S80, N98, R124, T135, T156, N158, Q159, T191, Q198, S273, Q277, S279, Q280, Q281, S283, N286, N315, T316, T319, S322, S346, E378, W386, N409, W410, N413. Residues near the GLUT3 pore entries are not chosen which are likely to be exposed to bulk water. b, Interaction frequency of polar/charged residues in N-domain interface with or without DMPE. The value in (a,b) is the average value from 0.6µs to 1.0µs. c, The average number of contacting water molecule to polar/charged residues in TMHs of GLUT3. ‘N’ and ‘C’ represent N, C-domain, respectively. Error bars represent SD (n = 4000 for each case). d, Interaction frequency of polar/charged residues in domain interfaces. The value in (c,d) is the average value from 0.6µs to 1.0µs. e, The average number of contacting water molecule to polar/charged residues in TMHs of GLUT3. Error bars represent SD (n = 2000 for each case). f, Interaction frequency of polar/charged residues in domain interfaces. The value in (e,f) is the average value from 0.3µs to 0.5µs.

Extended Data Fig. 9 Folding characteristics with PE lipid bicelles.

a, Representative folding trace of WT GLUT3 with PE-containing bicelle. Inset shows close-up view of the folding trace. The definition of each trace is identical to the traces in Extended Data Fig. 2d. Two replicates are shown. b, BIC values for the indicated number of states with 15 mol% PE bicelle (n = 11). The intermediate was largely preserved upon addition of DMPE lipids. c, Representative folding trace of WT GLUT3 with PE-containing bicelle in the presence of EMC. d, BIC values for the indicated number of states with 15 mol% PE bicelle in the presence of EMC (n = 10).

Extended Data Fig. 10 Insertion energy of sugar transporters.

a, Insertion energy histogram estimated for TMHs of all sugar porters. b, Insertion energy histogram estimated for C-domain TMHs of metazoan sugar porters. c, Insertion energy histogram estimated for C-domain TMHs of bacteria sugar porters. d, P-values from the Bartlett and Levene tests. 2 sets are used for statistical testing. e, Scatter plot of mean of top 3 insertion energy for N-domain as x-axis and mean of 3 top insertion energy for C-domain as y-axis for sugar transporters. f, Scatter plot of insertion energy variance for N-domain as x-axis and insertion energy variance for C-domain as y-axis. g, Average values of BLOSUM62 score for QLSQQLS motif is calculated for each group. (n = 26, 28, 54 and 24 for bacteria, metazoa, fungi and viridiplantae).

Supplementary information

Supplementary Information

Supplementary Figs 1−9.

Reporting Summary

Supplementary Data

Unprocessed western blot of Supplementary Fig. 7.

Source data

Source Data Extended Data Fig. 1

Unprocessed gels for Extended Data Fig. 1.

Source Data Extended Data Fig. 3

Unprocessed gels for Extended Data Fig. 3.

Source Data Extended Data Fig. 5

Unprocessed gels for Extended Data Fig. 5.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Choi, HK., Kang, H., Lee, C. et al. Evolutionary balance between foldability and functionality of a glucose transporter. Nat Chem Biol 18, 713–723 (2022). https://doi.org/10.1038/s41589-022-01002-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-022-01002-w

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