research communications\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL BIOLOGY
COMMUNICATIONS
ISSN: 2053-230X

Structural basis for oligomerization of the prokaryotic peptide transporter PepTSo2

CROSSMARK_Color_square_no_text.svg

aDepartment of Biological Sciences, Graduate School of Science, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan, bInstitute for Protein Research, Osaka University, 3-2 Yamadaoka, Suita, Osaka 565-0871, Japan, and cRIKEN Advanced Science Institute, 2-1 Hirosawa, Wako-shi, Saitama 351-0198, Japan
*Correspondence e-mail: nureki@bs.s.u-tokyo.ac.jp

Edited by I. Tanaka, Hokkaido University, Japan (Received 8 January 2019; accepted 13 March 2019; online 24 April 2019)

Proton-dependent oligopeptide transporters (POTs) belong to the major facilitator superfamily (MFS) and transport dipeptides and tripeptides from the extracellular environment into the target cell. The human POTs PepT1 and PepT2 are also involved in the absorption of various orally ingested drugs. Previously reported structures revealed that the bacterial POTs possess 14 helices, of which H1–H6 and H7–H12 constitute the typical MFS fold and the residual two helices are involved in the cytoplasmic linker. PepTSo2 from Shewanella oneidensis is a unique POT which reportedly assembles as a 200 kDa tetramer. Although the previously reported structures suggested the importance of H12 for tetramer formation, the structural basis for the PepTSo2-specific oligomerization remains unclear owing to the lack of a high-resolution tetrameric structure. In this study, the expression and purification conditions for tetrameric PepTSo2 were optimized. A single-particle cryo-EM analysis revealed the tetrameric structure of PepTSo2 incorporated into Salipro nanoparticles at 4.1 Å resolution. Furthermore, a combination of lipidic cubic phase (LCP) crystallization and an automated data-processing system for multiple microcrystals enabled crystal structures of PepTSo2 to be determined at resolutions of 3.5 and 3.9 Å. The present structures in a lipid bilayer revealed the detailed mechanism for the tetrameric assembly of PepTSo2, in which a characteristic extracellular loop (ECL) interacts with two asparagine residues on H12 which were reported to be important for tetramerization and plays an essential role in oligomeric assembly. This study provides valuable insights into the oligomerization mechanism of this MFS-type transporter, which will further pave the way for understanding other oligomeric membrane proteins.

1. Introduction and research aim

Proton-dependent oligopeptide transporters (POTs) belong to the major facilitator superfamily (MFS) and transport short-chain peptides from the extracellular environment into the target cell. The human POTs PepT1 and PepT2 are involved in the absorption of digested peptides and various orally ingested drugs, including antibiotics, antivirals and anticancer agents. Previously reported structures revealed that the bacterial POTs possess 14 helices, of which H1–H6 and H7–H12 constitute the typical MFS fold and the residual two helices, HA and HB, are inserted between H6 and H7 (Doki et al., 2013[Doki, S., Kato, H. E., Solcan, N., Iwaki, M., Koyama, M., Hattori, M., Iwase, N., Tsukazaki, T., Sugita, Y., Kandori, H., Newstead, S., Ishitani, R. & Nureki, O. (2013). Proc. Natl Acad. Sci. USA, 110, 11343-11348.]; Guettou et al., 2013[Guettou, F., Quistgaard, E. M., Trésaugues, L., Moberg, P., Jegerschöld, C., Zhu, L., Jong, A. J. O., Nordlund, P. & Löw, C. (2013). EMBO Rep. 14, 804-810.]; Newstead et al., 2011[Newstead, S., Drew, D., Cameron, A. D., Postis, V. L., Xia, X., Fowler, P. W., Ingram, J. C., Carpenter, E. P., Sansom, M. S., McPherson, M. J., Baldwin, S. A. & Iwata, S. (2011). EMBO J. 30, 417-426.]; Parker et al., 2017[Parker, J. L., Li, C., Brinth, A., Wang, Z., Vogeley, L., Solcan, N., Ledderboge-Vucinic, G., Swanson, J. M. J., Caffrey, M., Voth, G. A. & Newstead, S. (2017). Proc. Natl Acad. Sci. USA, 114, 13182-13187.]; Zhao et al., 2014[Zhao, Y., Mao, G., Liu, M., Zhang, L., Wang, X. & Zhang, X. C. (2014). Structure, 22, 1152-1160.]). PepTSo2 from Shewanella oneidensis is a unique POT transporter that has been reported to assemble as a 200 kDa tetramer in the detergent-solubilized form (Guettou et al., 2013[Guettou, F., Quistgaard, E. M., Trésaugues, L., Moberg, P., Jegerschöld, C., Zhu, L., Jong, A. J. O., Nordlund, P. & Löw, C. (2013). EMBO Rep. 14, 804-810.]). Two crystal structures of PepTSo2 have been reported: tetrameric and dimeric structures determined at resolutions of 4.6 and 3.2 Å, respectively (Guettou et al., 2013[Guettou, F., Quistgaard, E. M., Trésaugues, L., Moberg, P., Jegerschöld, C., Zhu, L., Jong, A. J. O., Nordlund, P. & Löw, C. (2013). EMBO Rep. 14, 804-810.]). Although the resolution of the dimeric structure was much higher, the dimer interface was stabilized by a zinc ion contained in the crystallization conditions, suggesting that this dimeric form represents a crystal-packing artifact. Recently, a cryo-electron microscopy (cryo-EM) structure of the PepTSo2 tetramer embedded in Salipro nanoparticles was reported at 6.5 Å resolution (Frauenfeld et al., 2016[Frauenfeld, J., Löving, R., Armache, J.-P., Sonnen, A. F.-P., Guettou, F., Moberg, P., Zhu, L., Jegerschöld, C., Flayhan, A., Briggs, J. A. G., Garoff, H., Löw, C., Cheng, Y. & Nordlund, P. (2016). Nature Methods, 13, 345-351.]). A structural comparison of the dimeric and tetrameric PepTSo2 structures suggested that subtle structural changes of H12 and the unconserved Asn468 on H12 are important for tetramer formation (Guettou et al., 2013[Guettou, F., Quistgaard, E. M., Trésaugues, L., Moberg, P., Jegerschöld, C., Zhu, L., Jong, A. J. O., Nordlund, P. & Löw, C. (2013). EMBO Rep. 14, 804-810.]). However, the structural basis for the PepTSo2-specific oligomerization remained unclear owing to the insufficient resolution of the tetrameric structure.

In this study, we overexpressed PepTSo2 in Escherichia coli and optimized the expression conditions. During the expression trials, we found that the number of histidine tags fused to the C-terminus drastically affected the oligomerization state of PepTSo2. Tetramer formation was confirmed by a single-particle analysis using cryo-EM and was further analyzed by X-ray crystallographic analyses using lipidic cubic phase (LCP) crystallization (Caffrey & Cherezov, 2009[Caffrey, M. & Cherezov, V. (2009). Nature Protoc. 4, 706-731.]). As a result of the crystallization trials, microcrystals were obtained under multiple conditions. We finally determined structures of PepTSo2 from two different crystal forms at 3.5 and 3.9 Å resolution by merging small wedge data sets from multiple microcrystals (Yamashita et al., 2018[Yamashita, K., Hirata, K. & Yamamoto, M. (2018). Acta Cryst. D74, 441-449.]).

PepTSo2 forms a tetramer in both of the crystal forms, and the structures superimpose well (with an r.m.s.d. of 0.83 Å for all Cα atoms). A slight movement of H12 was also observed, as in the previous structure, suggesting the importance of H12 for tetramer formation. Furthermore, the tetrameric assembly appears to be stabilized by hydrogen bonds between the asparagine residues in H12 and the unconserved extracellular loop (ECL), which was not well resolved in the previously determined low-resolution tetrameric structures (Frauenfeld et al., 2016[Frauenfeld, J., Löving, R., Armache, J.-P., Sonnen, A. F.-P., Guettou, F., Moberg, P., Zhu, L., Jegerschöld, C., Flayhan, A., Briggs, J. A. G., Garoff, H., Löw, C., Cheng, Y. & Nordlund, P. (2016). Nature Methods, 13, 345-351.]; Guettou et al., 2013[Guettou, F., Quistgaard, E. M., Trésaugues, L., Moberg, P., Jegerschöld, C., Zhu, L., Jong, A. J. O., Nordlund, P. & Löw, C. (2013). EMBO Rep. 14, 804-810.]). Although the physiological significance of the tetrameric assembly of PepTSo2 remains unclear, the oligomerization is possibly essential for the stable expression of PepTSo2 in the plasma membrane. This study provides valuable insights into the oligomerization mechanism of this MFS-type transporter, which will further facilitate the understanding of other oligomeric membrane proteins.

2. Materials and methods

2.1. Macromolecule production

The gene encoding full-length PepTSo2 (UniProt Q8EHE6) was subcloned into a pET-modified vector (Nishizawa et al., 2013[Nishizawa, T., Kita, S., Maturana, A. D., Furuya, N., Hirata, K., Kasuya, G., Ogasawara, S., Dohmae, N., Iwamoto, T., Ishitani, R. & Nureki, O. (2013). Science, 341, 168-172.]) which contains the DNA sequence encoding the Tobacco etch virus (TEV) protease cleavage site (ENLYFQG) followed by a His6 or His8 tag at the 3′-terminus. The macromolecule-production information is summarized in Table 1[link]. All of the PepTSo2 mutants were produced by a PCR-based site-directed mutagenesis method using PrimeSTAR Max DNA polymerase (Takara).

Table 1
Macromolecule-production information

Source organism S. oneidensis
DNA source UniProt Q8EHE6
Expression vector Modified pET-28a
Expression host E. coli Rosetta 2 (DE3)
Complete amino-acid sequence of the construct produced MTLGTNQVSKTHSFMTVSLIELWERFGYYGMQALIVYFMVQRLGFDDSRANLVWSACAALIYVSPAIGGWVGDKILGTKRTMLLGAGILSVGYALMTVPTENTWFMFSALGVIVVGNGLFKPNAGNLVRKIYEGDDSKIDSAFTIYYMAVNVGSTFSMLLTPWIKDYVNAQYGNEFGWHAAFAVCCVGILVGLGNYALMHKSLANYGSEPDTRPVNKKSLAIVLALAALSVVASAIILEYEDVARVFVYAAGVAVLGIFFHLIRTSEPSERAGLIAALILTVQTVFFFIFYQQMSTSLALFALRNVDWDFQVFGTHLWTWSPAQFQALNPIWIMVLSPVLAWSYSWAGRNNKDFSIAAKFALGFAVVAIGFFIYGFAGQFAVNGKTSSWVMIWGYASYSLGELLVSGLGLAMIARYVPARMGGFMMGAYFVASGISQYLGGVVANFASVPQDLVDPLQTLPVYTNLFNKLGVAAVVCTIIALAVLPLMRRLTESHHAHSSIENNAAASLRDVKAEQLESSGENLYFQ(GQFTSSVHHHHHH)
†The cloning artifacts are underlined. The residues cleaved by TEV protease are in parentheses.

2.2. Optimization of the expression conditions for PepTSo2

The plasmid was introduced into Escherichia coli Rosetta 2 (DE3) cells (Novagen) or genetically modified C41 (DE3) cells harboring the pRARE plasmids encoding the tRNAs for codons that are rarely used in E. coli. The transformed cells were grown in 2.5 ml Luria–Bertani (LB) medium containing 50 µg ml−1 ampicillin with or without 0.4% glucose or in 2.5 ml Terrific Broth (TB) at 310 K. When the absorbance at 600 nm (A600) reached 0.5–0.8, protein expression was induced with 0.5 mM isopropyl β-D-1-thio­galactopyranoside (IPTG) and the cells were grown for about 18 h at 293 K. The cells were centrifuged at 8000g for 1 min and were then suspended in solubilization buffer [50 mM Tris–HCl pH 8.0, 300 mM NaCl, 15 mM imidazole, 0.5 mM Tris(2-carboxyethyl)phos­phine (TCEP), 10% glycerol, 2% n-dodecyl-β-D-maltoside (DDM)] supplemented with cOmplete EDTA-free Protease Inhibitor Cocktail (Roche). The cells were disrupted by sonication using a Bioruptor UCW-310 (Cosmo Bio) and solubilized by gentle rotation at 277 K for 1.5 h. After the removal of unsolubilized materials by ultracentrifugation at 104 000g for 30 min at 277 K, the supernatant was mixed with the P3NTA probe (produced in-house; Backmark et al., 2013[Backmark, A. E., Olivier, N., Snijder, A., Gordon, E., Dekker, N. & Ferguson, A. D. (2013). Protein Sci. 22, 1124-1132.]) and analyzed by fluorescence-detection size-exclusion chromatography (FSEC; Kawate & Gouaux, 2006[Kawate, T. & Gouaux, E. (2006). Structure, 14, 673-681.]).

2.3. Large-scale expression and purification of PepTSo2

The plasmid containing C-terminally His6-tag-fused PepTSo2 was introduced into E. coli Rosetta 2 (DE3) cells (Novagen). The transformed cells were grown in 18 l LB medium containing 50 µg ml−1 ampicillin at 310 K. When the A600 reached 0.5–0.8, expression was induced with 0.5 mM IPTG and the cells were grown for about 18 h at 293 K. The cells were suspended in phosphate-buffered saline. The suspension was homogenized and mixed on ice, and was disrupted by 4–5 passes at 130 MPa using a Microfluidizer (Microfluidics). After centrifugation at 25 000g for 20 min to remove the debris, the supernatant was ultracentrifuged at 125 000g for 1 h at 277 K using a Type 45 Ti rotor (Beckman Coulter). The membrane fraction was resuspended in membrane buffer (50 mM Tris–HCl pH 8.0, 300 mM NaCl, 0.5 mM TCEP, 10% glycerol) supplemented with cOmplete EDTA-free Protease Inhibitor Cocktail (Roche) and stored at 193 K until use.

The membrane fraction was solubilized in solubilization buffer supplemented with cOmplete EDTA-free Protease Inhibitor Cocktail (Roche) at 277 K for 2 h. After removal of unsolubilized material by ultracentrifugation at 125 000g for 30 min at 277 K using a Type 45 Ti rotor (Beckman Coulter), the supernatant was mixed with 25 ml Ni–NTA Superflow resin (Qiagen) equilibrated with Ni buffer (50 mM Tris–HCl pH 8.0, 300 mM NaCl, 0.5 mM TCEP, 5% glycerol, 0.05% DDM) containing 15 mM imidazole for 1 h at 277 K. The mixture was transferred into an Econo-column (Bio-Rad) and the flowthrough fraction was collected. The resin was washed with ten column volumes of Ni buffer containing 50 mM imidazole, and the protein sample was eluted with Ni buffer containing 300 mM imidazole. To cleave the His6 tag that was fused to the C-terminus of PepTSo2, His-tagged TEV protease (produced in-house) was added to the eluted fraction to a 1:75(w:w) protease:protein ratio and the solution was dialyzed twice against 1 l dialysis buffer (50 mM Tris–HCl pH 8.0, 300 mM NaCl, 0.25 mM TCEP, 5% glycerol, 0.03% DDM). To remove the cleaved His tag and His-tagged TEV protease, the solution was mixed with Ni–NTA resin equilibrated with dialysis buffer for 1.5 h at 277 K and the collected flowthrough fraction was concentrated using an Amicon Ultra centrifugal filter (100 kDa molecular-weight cutoff; Millipore). The concentrated sample was ultracentrifuged at 40 000 rev min−1 for 20 min using an S55A2 rotor (Hitachi Koki) and the supernatant was applied onto a HiLoad 16/600 Superdex 200 pg column (GE Healthcare) equilibrated with SEC buffer consisting of 50 mM Tris–HCl pH 8.0, 300 mM NaCl, 5% glycerol, 0.03% DDM. The purity and homogeneity of the protein sample were assessed by FSEC. The purity of the protein sample was also assessed by SDS–PAGE. The protein concentrations were estimated based on the absorbance at 280 nm (A280) as measured using a NanoDrop spectrophoto­meter (Thermo Fisher Scientific). The protein bands detected in the SDS–PAGE gel were analyzed using the peptide mass fingerprint method.

2.4. Expression and purification of saposin A

The coding region of saposin A was cloned into a modified pET-28b vector that harbors an N-terminal His6 tag followed by the `eTEV' sequence (Tabata et al., 2018[Tabata, S., Kitago, Y., Fujii, Y., Mihara, E., Tamura-Kawakami, K., Norioka, N., Takahashi, K., Kaneko, M. K., Kato, Y. & Takagi, J. (2018). Protein Expr. Purif. 147, 94-99.]), resulting in the protein sequence MGSSHHHHHHSSGLVPRENLYFQGMGSLPCDICKDVVTAAGDMLKDNATEEEILVYLEKTCDWLPKPNMSASCKEIVDSYLPVILDIIKGEMSRPGEVCSALNLCES. The protein cleaved by TEV protease represents the saposin A polypeptide sequence with three additional N-terminal residues (Gly-Met-Gly). The expression and purification of saposin A were performed according to a previously reported method (Frauenfeld et al., 2016[Frauenfeld, J., Löving, R., Armache, J.-P., Sonnen, A. F.-P., Guettou, F., Moberg, P., Zhu, L., Jegerschöld, C., Flayhan, A., Briggs, J. A. G., Garoff, H., Löw, C., Cheng, Y. & Nordlund, P. (2016). Nature Methods, 13, 345-351.]).

2.5. Reconstruction of PepTSo2 into Salipro particles

For the reconstruction of PepTSo2 into Salipro nanoparticles, 8 µl of purified PepTSo2 at 20.4 mg ml−1 in TBSGD buffer (25 mM Tris–HCl pH 8.0, 5% glycerol, 0.03% DDM), 30 µl of a 5 mg ml−1 DMPG (Avanti) lipid solution in lipid-solubilization buffer (50 mM HEPES–NaOH pH 7.5, 150 mM NaCl, 0.224% DDM) and 12 µl HBSD buffer (50 mM HEPES–NaOH pH 7.5, 150 mM NaCl, 0.03% DDM) were mixed and incubated at 298 K for 5 min. Next, 120 µl of purified saposin A at 1.2 mg ml−1 was added and the mixture was incubated at 278 K for 5 min. Subsequently, 114 µl saposin SEC buffer, consisting of 25 mM sodium phosphate pH 7.4, 200 mM NaCl, was added to the mixture and incubated at 278 K for 5 min, after which 234 µl saposin SEC buffer was added. The mixture was injected into a Superdex 200 Increase 10/300 GL column (GE Healthcare) equilibrated with saposin SEC buffer. Fractions containing Salipro–PepTSo2 were pooled and concentrated to ∼0.5 mg ml−1 using an Amicon Ultra centrifugal filter (30 kDa molecular-weight cutoff; Millipore).

2.6. Cryo-EM sample preparation and data collection

Quantifoil copper 200 mesh R1.2/1.3 holey carbon grids were glow-discharged on a glass slide for 30 s. A 2.6 µl aliquot of the sample solution was applied onto the grid and blotted by filter paper for 4.5 s at 100% humidity and 277 K, and the grid was then quickly frozen by rapidly plunging it into liquid ethane using a Vitrobot Mark IV (Thermo Fisher Scientific). The grid was inserted into a Talos Arctica FEG transmission electron microscope (Thermo Fisher Scientific) operated at 200 kV, with the cryo-specimen stage cooled with liquid nitrogen. Cryo-EM images were recorded using a Falcon 3EC 4k × 4k CMOS direct electron detector (Thermo Fisher Scientific) in counting mode at a nominal magnification of ×96 000, corresponding to an image pixel size of 0.87 Å, using the EPU software package. Movie frames of the PepTSo2 tetramer embedded in Salipro were collected from 30° to 50° tilts of the stage in 10° increments at a dose rate of 1.2 e per pixel per second and an exposure time of 68.38 s. The total accumulated exposure of 82 e Å−2 was fractionated into 102 frames. A data set of 1609 micrographs (30° tilt, 1275 micrographs; 40° tilt, 272 micrographs; 50° tilt, 62 micrographs) was collected in a single session using a defocus range between 1.0 and 3.0 µm.

2.7. Image processing

Data sets from the various tilts from 30° to 50° were used. The movie frames were subsequently aligned to correct for beam-induced movement and drift using MotionCor2 (Zheng et al., 2017[Zheng, S. Q., Palovcak, E., Armache, J., Verba, K. A., Cheng, Y. & Agard, D. A. (2017). Nature Methods, 14, 331-332.]), and the parameters for the contrast transfer function (CTF) were estimated using Gctf (Zhang, 2016[Zhang, K. (2016). J. Struct. Biol. 193, 1-12.]). A total of 288 417 particle images were automatically picked from 1484 micrographs using Gautomatch (https://www.mrc-lmb.cam.ac.uk/kzhang/), and two-dimensional (2D) and three-dimensional (3D) classifications were then performed using Relion-3.0 (Zivanov et al., 2018[Zivanov, J., Nakane, T., Forsberg, B. O., Kimanius, D., Hagen, W. J. H., Lindahl, E. & Scheres, S. H. W. (2018). Elife, 7, e42166.]). Particle images from the good 2D class were selected to obtain the initial 3D model of the PepTSo2 tetramer embedded in Salipro using cryoSPARC (Punjani et al., 2017[Punjani, A., Rubinstein, J. L., Fleet, D. J. & Brubaker, M. A. (2017). Nature Methods, 14, 290-296.]) with C4 symmetry. A total of 43 172 particles from the best 3D class were subjected to 3D refinement, which produced a reconstruction with a resolution of 5.9 Å and a B factor of −307 Å2. The 3D-refined structure was further refined with per-particle defocus with a search range of ±2000 Å and Bayesian polishing, which improved the resolution to 4.3 Å with a B factor of −171 Å2. To improve the resolution further, CTF refinement with beam-tilt correction and Bayesian polishing was iterated two times, giving 4.3 Å resolution and a B factor of −161 Å2 and finally 4.1 Å resolution and a B factor of −147 Å2. The 3D density map was visualized using UCSF Chimera (Pettersen et al., 2004[Pettersen, E. F., Goddard, T. D., Huang, C. C., Couch, G. S., Greenblatt, D. M., Meng, E. C. & Ferrin, T. E. (2004). J. Comput. Chem. 25, 1605-1612.]).

2.8. X-ray crystallographic analyses

The PepTSo2 sample was mixed with monoolein (Nu-Chek Prep) at a 2:3(w:w) protein:lipid ratio using coupled syringes. Drops of the mixture (30 nl) were dispensed onto 96-well Laminex glass sandwich plates (MD11-50-100; Molecular Dimensions) and overlaid with 800 nl reservoir solution using a Gryphon LCP crystallization robot (Art Robbins). The initial crystallization screening was performed at 293 K using in-house-produced grid-screening crystallization kits as reservoir solutions. To optimize the crystallization conditions, StockOptions Salt and Additive Screen (Hampton Research) was added to the reservoir solutions for LCP crystallization, in addition to optimization of the pH and the concentrations of precipitants and salts. Crystals were picked up using MicroMeshes (MiTeGen), MicroMounts (MiTeGen) or LithoLoops (Protein Wave) and were flash-cooled in liquid nitrogen. Crystallization information is provided in Table 2[link].

Table 2
Crystallization conditions

  Form A crystal Form B crystal
Method Lipidic cubic phase (LCP) Lipidic cubic phase (LCP)
Plate type 96-well glass sandwich plate 96-well glass sandwich plate
Temperature (K) 293 293
Protein concentration (mg ml−1) 45 45
Buffer composition of protein solution 300 mM NaCl, 50 mM Tris–HCl pH 8.0, 5% glycerol, 0.03% DDM 300 mM NaCl, 50 mM Tris–HCl pH 8.0, 5% glycerol, 0.03% DDM
Composition of reservoir solution 40% PEG 200, 100 mM sodium malonate pH 7.0, 100 mM sodium acetate pH 4.0 40% PEG 200, 100 mM NaCl, 100 mM sodium acetate pH 4.0
Volume of LCP drop (nl) 30 30
Volume of reservoir (nl) 800 800

2.9. X-ray data collection and processing

All X-ray diffraction data sets were collected by the helical data-collection method using the microfocus beam at SPring-8 beamline BL32XU (Hirata et al., 2013[Hirata, K., Kawano, Y., Ueno, G., Hashimoto, K., Murakami, H., Hasegawa, K., Hikima, T., Kumasaka, T. & Yamamoto, M. (2013). J. Phys. Conf. Ser. 425, 012002.]). The locations of well diffracting crystals were identified by raster scanning, and a 5° or 10° wedge of data was collected from each crystal. All diffraction data were processed with XDS (Kabsch, 2010[Kabsch, W. (2010). Acta Cryst. D66, 125-132.]) and were merged with XSCALE based on hierarchical clustering analysis with the cross-correlation method implemented in the KAMO software (Yamashita et al., 2018[Yamashita, K., Hirata, K. & Yamamoto, M. (2018). Acta Cryst. D74, 441-449.]).

Molecular replacement was performed with Phaser (McCoy et al., 2007[McCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C. & Read, R. J. (2007). J. Appl. Cryst. 40, 658-674.]). For structure determination of the form B crystal, the structure of molecule A of PepTSo2 (PDB entry 4lep), reported as a dimeric form, was used as the search model. For structure determination of the form A crystal, the structure determined from the form B crystal (PDB entry 6jkc) was used as the search model. Model building was performed with Coot (Emsley & Cowtan, 2004[Emsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126-2132.]; Emsley et al., 2010[Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. (2010). Acta Cryst. D66, 486-501.]). Refinement was performed with phenix.refine (Adams et al., 2010[Adams, P. D., Afonine, P. V., Bunkóczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd, J. J., Hung, L.-W., Kapral, G. J., Grosse-Kunstleve, R. W., McCoy, A. J., Moriarty, N. W., Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C. & Zwart, P. H. (2010). Acta Cryst. D66, 213-221.]) and REFMAC5 (Murshudov et al., 2011[Murshudov, G. N., Skubák, P., Lebedev, A. A., Pannu, N. S., Steiner, R. A., Nicholls, R. A., Winn, M. D., Long, F. & Vagin, A. A. (2011). Acta Cryst. D67, 355-367.]) using the CCP4i2 interface (Potterton et al., 2018[Potterton, L., Agirre, J., Ballard, C., Cowtan, K., Dodson, E., Evans, P. R., Jenkins, H. T., Keegan, R., Krissinel, E., Stevenson, K., Lebedev, A., McNicholas, S. J., Nicholls, R. A., Noble, M., Pannu, N. S., Roth, C., Sheldrick, G., Skubak, P., Turkenburg, J., Uski, V., von Delft, F., Waterman, D., Wilson, K., Winn, M. & Wojdyr, M. (2018). Acta Cryst. D74, 68-84.]). The data-collection statistics are summarized in Table 3[link]. Molecular graphics were generated using CueMol (https://www.cuemol.org).

Table 3
X-ray data-collection and refinement statistics

Values in parentheses are for the outer shell.

  Form A crystal (PDB entry 6jkd) Form B crystal (PDB entry 6jkc)
X-ray data collection
 Diffraction source BL32XU, SPring-8 BL32XU, SPring-8
 Wavelength (Å) 1.0000 1.0000
 Temperature (K) 100 100
 Detector EIGER X 9M, Dectris EIGER X 9M, Dectris
 Crystal-to-detector distance (mm) 300 280
 Rotation range per image (°) 0.1 0.1
 Rotation range per crystal (°) 10 5
 No. of crystals 10 9
 Space group I4 P4212
a, b, c (Å) 115.1, 115.1, 110.1 119.4, 119.4, 104.3
 Resolution range (Å) 50–3.90 (4.14–3.90) 47.53–3.50 (3.71–3.50)
 Total No. of reflections 23358 29586
 No. of unique reflections 6448 8312
 Completeness (%) 97.4 (97.8) 83.0 (83.4)
 Multiplicity 3.6 (3.7) 3.6 (3.5)
 〈I/σ(I)〉 4.38 (1.44) 5.49 (1.22)
Rmeas 0.455 (1.277) 0.328 (1.586)
 CC1/2 0.946 (0.459) 0.961 (0.293)
Refinement
 Resolution (Å) 46.65–3.90 47.53–3.50
Rwork/Rfree 0.2524/0.2832 0.2672/0.3030
 No. of atoms
  Protein 3476 3498
  Ligand 0 0
  Solvent 0 0
 Average B factors (Å2)
  Protein 70.3 65.7
  Ligand
  Solvent
 R.m.s. deviations
  Bond lengths (Å) 0.002 0.003
  Bond angles (°) 0.51 0.55
 Ramachandran plot
  Favored (%) 97.1 96.9
  Allowed (%) 2.9 3.1
  Outliers (%) 0 0

2.10. Model building for cryo-EM

The crystal structure of PepTSo2 determined in this work (form B crystal) was fitted as a rigid body into the 3D density map using UCSF Chimera and Coot (Emsley & Cowtan, 2004[Emsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126-2132.]; Emsley et al., 2010[Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. (2010). Acta Cryst. D66, 486-501.]). Refinement was performed with phenix.refine (Adams et al., 2010[Adams, P. D., Afonine, P. V., Bunkóczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd, J. J., Hung, L.-W., Kapral, G. J., Grosse-Kunstleve, R. W., McCoy, A. J., Moriarty, N. W., Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C. & Zwart, P. H. (2010). Acta Cryst. D66, 213-221.]). The conformations were refined using real-space refinement in PHENIX (Adams et al., 2010[Adams, P. D., Afonine, P. V., Bunkóczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd, J. J., Hung, L.-W., Kapral, G. J., Grosse-Kunstleve, R. W., McCoy, A. J., Moriarty, N. W., Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C. & Zwart, P. H. (2010). Acta Cryst. D66, 213-221.]) with secondary-structure restraints. We first performed model refinement for each subunit separately against the corresponding EM maps. To resolve the possible clashes in the subunit interfaces, we refined the entire models against the corresponding EM maps. Table 4[link] summarizes the refinement statistics for the overall structure, the deposited maps and their associated coordinates.

Table 4
Cryo-EM data-collection and refinement statistics

  PepTSo2 embedded in Salipro nanoparticles (PDB entry 6ji1, EMDB ID EMD-9832)
  0° tilt 30° tilt 40° tilt 50° tilt
Cryo-EM data collection
 TEM Titan Krios Talos Arctica
 Accelerating voltage (kV) 300 200
 Camera Falcon III (counting) Falcon III (counting)
 Total dose (e Å−2) 60 82
 No. of micrographs 864 1275 272 62
 No. of particles 130487 229633 49304 9480
3D refinement
 Resolution (Å) 6.7 4.13
 Map-sharpening B factor (Å2) −200 −147
 Fourier shell correlation criterion 0.143 0.143
 Particles used in final 3D refinement 3415 43172
Coordinate refinement and validation
 R.m.s. deviations
  Bond lengths (Å)   0.010
  Bond angles (°)   1.225
 Ramachandran plot
  Favored (%)   83.63
  Allowed (%)   11.06
  Outliers (%)   5.31
MolProbity score   2.09
 Clashscore (all-atom)   7.36

3. Results

3.1. FSEC-based optimization of expression conditions

For structural analysis of tetrameric PepTSo2, the expression and purification conditions were optimized. Full-length PepTSo2 fused with a His tag at the C-terminus was expressed using a modified pET vector (Nishizawa et al., 2013[Nishizawa, T., Kita, S., Maturana, A. D., Furuya, N., Hirata, K., Kasuya, G., Ogasawara, S., Dohmae, N., Iwamoto, T., Ishitani, R. & Nureki, O. (2013). Science, 341, 168-172.]; Table 1[link]), and the expression level and homogeneity of the samples were evaluated by an FSEC-based high-throughput screening method (Kawate & Gouaux, 2006[Kawate, T. & Gouaux, E. (2006). Structure, 14, 673-681.]) using the P3NTA peptide as a probe (Backmark et al., 2013[Backmark, A. E., Olivier, N., Snijder, A., Gordon, E., Dekker, N. & Ferguson, A. D. (2013). Protein Sci. 22, 1124-1132.]). Unexpectedly, the number of histidine residues fused to the C-terminus drastically affected the oligomerization state of PepTSo2, and the His6-tagged PepTSo2 formed a tetramer more readily (Fig. 1[link]a). Under the conditions that we tested, the expression levels of PepTSo2 were greater in the BL21 (DE3) derivative strain of E. coli compared with the C41 (DE3) strain, which was used in the previous studies (Guettou et al., 2013[Guettou, F., Quistgaard, E. M., Trésaugues, L., Moberg, P., Jegerschöld, C., Zhu, L., Jong, A. J. O., Nordlund, P. & Löw, C. (2013). EMBO Rep. 14, 804-810.]). The presence or absence of tRNAs reading rare codons in the expression host did not affect the expression level. The addition of glucose to the medium did not affect the expression level. The expression level of PepTSo2 was higher when using LB medium compared with the TB medium used in the previous study. These culture conditions affected the expression level but not the oligomerization state of PepTSo2.

[Figure 1]
Figure 1
Protein preparation. (a) Fluorescence-detection size-exclusion chromatography (FSEC)-based analysis of His6 (black) and His8 (red) constructs. The histidine tag-specific fluorescent probe P3NTA was used for detection (excitation, 482 nm; emission, 520 nm). The peak corresponding to tetrameric PepTSo2 is marked with an asterisk. (b) SEC chromatogram of purified PepTSo2. The fluorescent signals, which were mainly derived from the tryptophan residues, were monitored by the fluorescence detector, with excitation at 280 nm and emission at 350 nm. The peak corresponding to tetrameric PepTSo2 is marked with an asterisk. (c) SDS–PAGE analysis with Coomassie Brilliant Blue staining. Left lane, molecular-weight markers (labeled in kDa); right lane, the merged peak fraction from the SEC purification.

3.2. Large-scale expression and purification of PepTSo2

C-terminally His6-tagged PepTSo2 was expressed on a large scale under the conditions optimized in this study. PepTSo2 solubilized with DDM was purified by Ni–NTA affinity chromatography followed by cleavage of the His tag, a second Ni–NTA chromatography step to remove the His-tagged TEV protease and size-exclusion chromatography (SEC). The SEC chromatogram peak showed monodispersity, and the peak fractions showed a single band in the SDS–PAGE analysis, demonstrating high purity and homogeneity of the obtained PepTSo2 sample (Figs. 1[link]b and 1[link]c). The elution volume of PepTSo2 in the SEC chromatogram was 14.8 ml in DDM-containing buffer using a Superose 6 Increase column (GE Healthcare) (Fig. 1[link]b), which was a reasonable elution position for tetrameric PepTSo2 (the molecular weight is about 200 kDa). The final yield was about 2.5 mg per litre of E. coli culture.

3.3. Single-particle cryo-EM analyses of PepTSo2

To gain structural insight into oligomer formation by PepTSo2, we reconstructed PepTSo2 into Salipro nanoparticles using saposin A and DMPG lipids, and performed single-particle analysis using cryo-electron microscopy (Figs. 2[link]a–2[link]e, Supplementary Fig. S1). Although square-shaped particles were clearly observed in the cryo-EM micrographs of PepTSo2 reconstructed in Salipro, the overall resolution was limited to 6.7 Å and the quality of the map was too poor to refine the atomic model owing to severe orientation bias (Supplementary Fig. S1). Interestingly, this preferred specimen-orientation problem was not reported in the previous study (Frauenfeld et al., 2016[Frauenfeld, J., Löving, R., Armache, J.-P., Sonnen, A. F.-P., Guettou, F., Moberg, P., Zhu, L., Jegerschöld, C., Flayhan, A., Briggs, J. A. G., Garoff, H., Löw, C., Cheng, Y. & Nordlund, P. (2016). Nature Methods, 13, 345-351.]). In the previous study, bovine brain lipid extract was used for reconstruction, but this product is no longer commercially accessible. The difference in the lipids used for reconstruction in this study might affect the physical properties of the specimen under solution conditions and thus cause the preferred specimen-orientation problem. To alleviate the orientation bias, we tried to collect data by tilting the grid from 30° to 50° (Tan et al., 2017[Tan, Y. Z., Baldwin, P. R., Davis, J. H., Williamson, J. R., Potter, C. S., Carragher, B. & Lyumkis, D. (2017). Nature Methods, 14, 793-796.]), which drastically improved the data quality (Supplementary Fig. S1). A total of 43 172 selected particles from three different tilt angles (30°, 40° and 50°) yielded a 3D EM map at an overall resolution of 4.1 Å, according to the gold-standard Fourier shell correlation (FSC) 0.143 criterion (Scheres & Chen, 2012[Scheres, S. H. W. & Chen, S. (2012). Nature Methods, 9, 853-854.]; Fig. 2[link]d, Supplementary Fig. S2, Table 4[link]). The tetrameric organization of PepTSo2 is clearly visualized in the final 3D density map (Fig. 2[link]f), in which the secondary-structural features and the side-chain densities from bulky amino-acid residues are clearer compared with the previous study (6.5 Å resolution; Frauenfeld et al., 2016[Frauenfeld, J., Löving, R., Armache, J.-P., Sonnen, A. F.-P., Guettou, F., Moberg, P., Zhu, L., Jegerschöld, C., Flayhan, A., Briggs, J. A. G., Garoff, H., Löw, C., Cheng, Y. & Nordlund, P. (2016). Nature Methods, 13, 345-351.]; Fig. 2[link]g). Residual density derived from the Salipro scaffold composed of saposin A and lipids was also observed around the tetrameric PepTSo2 molecule; however, the density in these regions was not so clear compared with the PepTSo2 molecule (Fig. 2[link]h), suggesting flexibility of the Salipro scaffold region. The EM density map suggested a contribution of the extracellular loops between the TM helices that were not well resolved in the previous study to the tetramer assembly (Figs. 2[link]h and 2[link]f). However, the atomic-level oligomerization mechanism remained unclear because of the resolution limit.

[Figure 2]
Figure 2
Single-particle analysis of PepTSo2 using cryo-electron microscopy. (a) Salipro reconstitution of PepTSo2. SEC chromatograms before and after reconstitution of PepTSo2 are shown as gray and black lines, respectively (left panel). The peak fractions indicated by the red arrow were analyzed by SDS–PAGE (right panel). (b) Raw electron micrograph of PepTSo2. The Salipro-reconstituted PepTSo2 particles are indicated by red circles. (c) Representative 2D class averages generated from the particles. The scale bar represents 10 nm. (d) Gold-standard FSC between two independently refined half-maps. (e) Euler angle distribution of all particles collected by the tilting method. (f) 3D density maps of the tetrameric PepTSo2 incorporated in the Salipro nanoparticle. The atomic model was refined by using the crystal structure determined in this study as the initial model. (g) Densities of selected helices and side chains. (h) The density around the inter-protomer interaction area. The extracellular loop of a protomer is colored green and H12 of the neighboring protomer is colored magenta.

3.4. X-ray crystallographic analyses of PepTSo2

To further improve the resolution, we tried to crystallize tetrameric PepTSo2 using the LCP method (Caffrey & Cherezov, 2009[Caffrey, M. & Cherezov, V. (2009). Nature Protoc. 4, 706-731.]) at protein concentrations of 20, 45 and 75 mg ml−1. Microcrystals of PepTSo2 were obtained using a 45 mg ml−1 sample under multiple conditions containing PEG 200 as a precipitant (Fig. 3[link]). Since the size of the crystals was too small and data collection from a single crystal was impossible, small wedge data were collected from microcrystals using the microfocus beam at BL32XU at SPring-8, Hyogo, Japan and were combined using the multi-crystal merging software KAMO (Yamashita et al., 2018[Yamashita, K., Hirata, K. & Yamamoto, M. (2018). Acta Cryst. D74, 441-449.]). By using the molecular-replacement method with the previously reported structure of the PepTSo2 dimer (PDB entry 4lep, molecule A) as the search model (Guettou et al., 2014[Guettou, F., Quistgaard, E. M., Raba, M., Moberg, P., Löw, C. & Nordlund, P. (2014). Nature Struct. Mol. Biol. 21, 728-731.]), the structures of the form A and B crystals were finally determined at resolutions of 3.9 and 3.5 Å, respectively (Table 3[link]).

[Figure 3]
Figure 3
LCP crystals of PepTSo2. (a) Crystals of form A. (b) Crystals of form B.

3.5. Overall structure of tetrameric PepTSo2

The PepTSo2 structures obtained from the two crystal forms (forms A and B) were assembled as tetramers in similar manners (r.m.s.d. of 0.83 Å for all Cα atoms), although the space groups and the crystal packings were different (Table 3[link], Supplementary Fig. S3). Thus, in the following we focus on the structure obtained from the form B crystal (Figs. 4[link]a, 4[link]b and 4[link]c), as the quality of its electron density is better than that of the other crystal form. All of the protomers in the tetramer adopted the inward-open conformation, in which a large central hydrophilic substrate-binding pocket faces the intracellular side of the membrane. The structural comparison shows that the present tetrameric structure superimposes well with the previously reported dimeric structure at the protomer level (r.m.s.d. of 1.1 Å for all Cα atoms; Supplementary Fig. S4a). The present crystal structures fitted well to the cryo-EM map obtained in this study (Fig. 2[link]f, Supplementary Fig. 4b).

[Figure 4]
Figure 4
Tetrameric organization of PepTSo2. (a, b, c) The overall structure of tetrameric PepTSo2 is viewed from the (a) extracellular, (b) membrane and (c) intracellular sides. The four protomers are colored magenta, green, blue and yellow. The rectangles labeled `d' and `e' indicate the regions highlighted in (d) and (e), respectively. (d, e) The intermolecular interactions of the extracellular loop (ECL) with H12 of the neighboring protomer are viewed from the (d) extracellular and (e) membrane sides. The possible hydrogen bonds of up to 3.2 Å are shown as black dotted lines. (f) A 2mFoDFc electron-density map of the interacting area between the ECL and H12 contoured at 1.1σ. The C atoms of the two protomers are colored magenta and green. (g) Sequence alignment of PepTSo2 with other prokaryotic POT families in the ECL area. (h) FSEC-based mutational analyses of PepTSo2. The peak corresponding to tetrameric PepTSo2 is marked with an asterisk.

3.6. Intermolecular interactions of the PepTSo2 tetramer

In the structure determined in this study, a sliding motion of H12 was also observed compared with the previously reported dimeric structure (Supplementary Fig. 4a), supporting the importance of H12 for tetramer formation suggested in the previous report (Guettou et al., 2014[Guettou, F., Quistgaard, E. M., Raba, M., Moberg, P., Löw, C. & Nordlund, P. (2014). Nature Struct. Mol. Biol. 21, 728-731.]). On the intracellular side, inter-subunit hydrophobic interactions were formed between H12 of one protomer and H5 and H8 of the neighboring protomer (Fig. 4[link]c). On the extracellular side of the present structure, electron density for the extracellular loop (ECL), which was not well resolved in the previous structures, was clearly observed (Fig. 4[link]f). The ECL consists of two extracellular linkers between H7–H8 (ECLa) and H9–H10 (ECLb) (Figs. 4[link]d and 4[link]e), in which four β-strands form parallel and antiparallel β-sheet structures. The structure suggested that the ECL plays an important role in PepTSo2 tetramer formation, together with H12 (Figs. 4[link]d and 4[link]e). The ECL interacts with H9 and H12 of the neighboring protomer by hydrophobic or hydrogen-bond interactions (Figs. 4[link]a, 4[link]d and 4[link]e). In the previous study, the importance of Asn468 in H12 for the tetramerization of PepTSo2 was suggested (Guettou et al., 2014[Guettou, F., Quistgaard, E. M., Raba, M., Moberg, P., Löw, C. & Nordlund, P. (2014). Nature Struct. Mol. Biol. 21, 728-731.]). The higher resolution structure determined in this study clearly shows that the side chains of Asn465 and Asn468 in H12 form hydrogen bonds to the ECL of the neighboring protomer, thus stabilizing the tetramer (Figs. 4[link]a, 4[link]d and 4[link]e). These two asparagine residues and the amino-acid sequence of the ECL are not conserved among the POT family members. Interestingly, the ECL of PepTSo2 is longer compared with those of other prokaryotic orthologues with structures that were reported to be monomers or dimers (Fig. 4[link]g). In contrast, among the POT family members that are phylogenetically closer to PepTSo2 at the amino-acid level, the peptide lengths of the ECL are preserved, while the amino-acid sequences are not highly conserved. These results suggest that these non­conserved amino-acid residues play an important role in PepTSo2-specific tetramer formation, which is further supported by the mutational analyses using FSEC (Fig. 4[link]h).

4. Discussion

Many membrane proteins exist and function as oligomers in the membrane. The oligomeric states are highly diverse among the individual membrane proteins, and are sometimes in equilibrium depending on the protein concentration in the membrane. For membrane transporters, it has been suggested that oligomerization plays various roles, such as in transport activity, membrane trafficking, regulation of function and turnover (Alguel et al., 2016[Alguel, Y., Cameron, A. D., Diallinas, G. & Byrne, B. (2016). Biochem. Soc. Trans. 44, 1737-1744.]). Oligomerization can allow transporters to form a stable functional structure. In some cases, such as ATP-binding cassette (ABC) transporters (Locher, 2016[Locher, K. P. (2016). Nature Struct. Mol. Biol. 23, 487-493.]) and the small multidrug transporter EmrE (Chen et al., 2007[Chen, Y.-J., Pornillos, O., Lieu, S., Ma, C., Chen, A. P. & Chang, G. (2007). Proc. Natl Acad. Sci. USA, 104, 18999-19004.]), the interface of the protomers forms the substrate-binding site and translocation pathway. Another example is the well studied bacterial glutamine transporter Gltph, which assembles as a trimer regardless of concentration, although each protomer contains the substrate-transport pathway (Yernool et al., 2004[Yernool, D., Boudker, O., Jin, Y. & Gouaux, E. (2004). Nature (London), 431, 811-818.]). Previous functional and structural analyses have suggested that the trimeric arrangement of Gltph plays a critical role in conformational changes during the transport cycle (Reyes et al., 2009[Reyes, N., Ginter, C. & Boudker, O. (2009). Nature (London), 462, 880-885.]). In the above cases, the manner of oligomeric arrangement is widely preserved at the family level and is critical for proper function.

In contrast, there are some membrane transporters that form species-specific or subtype-specific oligomers, in which each protomer consists of a transport unit. This study provides structural insight into one of these types of transporters, PepTSo2, which belongs to the POT family of the MFS. PepTSo2 is derived from the Gram-negative anaerobic bacterium S. oneidensis and is known to assemble as a tetramer (Guettou et al., 2014[Guettou, F., Quistgaard, E. M., Raba, M., Moberg, P., Löw, C. & Nordlund, P. (2014). Nature Struct. Mol. Biol. 21, 728-731.]). In contrast, another POT family protein from S. oneidensis, PepTSo, exists as a monomer in solution (Guettou et al., 2014[Guettou, F., Quistgaard, E. M., Raba, M., Moberg, P., Löw, C. & Nordlund, P. (2014). Nature Struct. Mol. Biol. 21, 728-731.]). Previous reports also revealed that other POT family members exist as mixtures of monomers and dimers, while tetramer formation has only been observed for PepTSo2 (Guettou et al., 2014[Guettou, F., Quistgaard, E. M., Raba, M., Moberg, P., Löw, C. & Nordlund, P. (2014). Nature Struct. Mol. Biol. 21, 728-731.]; Löw et al., 2013[Löw, C., Moberg, P., Quistgaard, E. M., Hedrén, M., Guettou, F., Frauenfeld, J., Haneskog, L. & Nordlund, P. (2013). Biochim. Biophys. Acta, 1830, 3497-3508.]). In this study, a combination of X-ray crystallographic and single-particle cryo-EM analyses revealed the species-specific tetramer formation mechanism of PepTSo2, in which the two asparagine residues in H12 and the unpreserved ECL play key roles. Although the physiological significance of tetramer formation remains unknown, tetramer formation may contribute to stable expression and/or membrane local­ization. In all of the structures determined in this study, each PepTSo2 protomer adopts a similar inward-open state, which might reflect cooperative substrate transport in the tetramer. It is also possible that the oligomerization state of PepTSo2 under physiological conditions is regulated by lipid molecules in the membrane, an unknown partner protein or the binding of substrates or additional cofactors. However, the molecular basis and the physiological significance of the oligomerization of membrane transporters, including PepTSo2, remain unclear at many points and further research is awaited.

5. Data availability

Data supporting the findings of this manuscript are available from the corresponding authors upon reasonable request. The coordinates and structure factors from the X-ray crystallo­graphic analyses have been deposited in the Protein Data Bank (PDB) with accession codes 6jkd and 6jkc for the form A (space group I4) and form B (space group P4212) crystals, respectively. X-ray diffraction images are also available from the Zenodo data repository (https://zenodo.org/record/2533841). The coordinates used for electron-microscopic analysis have been deposited in the PDB with accession code 6ji1. The EM map of PepTSo2 embedded in Salipro nanoparticles has been deposited in the Electron Microscopy Data Bank (EMDB) under accession code EMD-9832.

Footnotes

These authors contributed equally to this work.

§Present address: Faculty of Pharmacy, Keio University, 1-5-30 Shibakoen, Minato-ku, Tokyo 105-8512, Japan.

Acknowledgements

We thank T. Nishizawa and T. Kusakizako for useful suggestions and discussions, and the beamline staff members (K. Hirata, N. Kawano and K. Yamashita) at BL32XU at SPring-8 for technical support during data collection. We also thank K. Iwasaki for help at the IPR cryo-EM facility. The X-ray diffraction experiments were performed on BL32XU at SPring-8 (proposal No. 2017B2578) with the approval of RIKEN.

Funding information

The following funding is acknowledged: Japan Society for the Promotion of Science (grant No. 16H06294). This work was partially supported by the Platform Project for Supporting Drug Discovery and Life Science Research (Basis for Supporting Innovative Drug Discovery and Life Science Research; BINDS) from AMED under grant Nos. JP18am0101072 and JP18am0101075, by a Grant-in-Aid for Scientific Research (S) (24227004) and a Grant-in-Aid for Scientific Research (B) (25291011) from the Japan Society for the Promotion of Science (JSPS) to ON and RI, respectively, and by a Grant-in-Aid for JSPS Fellows.

References

First citationAdams, P. D., Afonine, P. V., Bunkóczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd, J. J., Hung, L.-W., Kapral, G. J., Grosse-Kunstleve, R. W., McCoy, A. J., Moriarty, N. W., Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C. & Zwart, P. H. (2010). Acta Cryst. D66, 213–221.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationAlguel, Y., Cameron, A. D., Diallinas, G. & Byrne, B. (2016). Biochem. Soc. Trans. 44, 1737–1744.  CrossRef CAS PubMed Google Scholar
First citationBackmark, A. E., Olivier, N., Snijder, A., Gordon, E., Dekker, N. & Ferguson, A. D. (2013). Protein Sci. 22, 1124–1132.  CrossRef CAS PubMed Google Scholar
First citationCaffrey, M. & Cherezov, V. (2009). Nature Protoc. 4, 706–731.  Web of Science CrossRef CAS Google Scholar
First citationChen, Y.-J., Pornillos, O., Lieu, S., Ma, C., Chen, A. P. & Chang, G. (2007). Proc. Natl Acad. Sci. USA, 104, 18999–19004.  CrossRef PubMed CAS Google Scholar
First citationDoki, S., Kato, H. E., Solcan, N., Iwaki, M., Koyama, M., Hattori, M., Iwase, N., Tsukazaki, T., Sugita, Y., Kandori, H., Newstead, S., Ishitani, R. & Nureki, O. (2013). Proc. Natl Acad. Sci. USA, 110, 11343–11348.  Web of Science CrossRef CAS PubMed Google Scholar
First citationEmsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126–2132.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEmsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. (2010). Acta Cryst. D66, 486–501.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFrauenfeld, J., Löving, R., Armache, J.-P., Sonnen, A. F.-P., Guettou, F., Moberg, P., Zhu, L., Jegerschöld, C., Flayhan, A., Briggs, J. A. G., Garoff, H., Löw, C., Cheng, Y. & Nordlund, P. (2016). Nature Methods, 13, 345–351.  CrossRef PubMed Google Scholar
First citationGuettou, F., Quistgaard, E. M., Raba, M., Moberg, P., Löw, C. & Nordlund, P. (2014). Nature Struct. Mol. Biol. 21, 728–731.  CrossRef CAS Google Scholar
First citationGuettou, F., Quistgaard, E. M., Trésaugues, L., Moberg, P., Jegerschöld, C., Zhu, L., Jong, A. J. O., Nordlund, P. & Löw, C. (2013). EMBO Rep. 14, 804–810.  CrossRef CAS PubMed Google Scholar
First citationHirata, K., Kawano, Y., Ueno, G., Hashimoto, K., Murakami, H., Hasegawa, K., Hikima, T., Kumasaka, T. & Yamamoto, M. (2013). J. Phys. Conf. Ser. 425, 012002.  CrossRef Google Scholar
First citationKabsch, W. (2010). Acta Cryst. D66, 125–132.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKawate, T. & Gouaux, E. (2006). Structure, 14, 673–681.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLocher, K. P. (2016). Nature Struct. Mol. Biol. 23, 487–493.  Web of Science CrossRef CAS Google Scholar
First citationLöw, C., Moberg, P., Quistgaard, E. M., Hedrén, M., Guettou, F., Frauenfeld, J., Haneskog, L. & Nordlund, P. (2013). Biochim. Biophys. Acta, 1830, 3497–3508.  PubMed Google Scholar
First citationMcCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C. & Read, R. J. (2007). J. Appl. Cryst. 40, 658–674.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMurshudov, G. N., Skubák, P., Lebedev, A. A., Pannu, N. S., Steiner, R. A., Nicholls, R. A., Winn, M. D., Long, F. & Vagin, A. A. (2011). Acta Cryst. D67, 355–367.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationNewstead, S., Drew, D., Cameron, A. D., Postis, V. L., Xia, X., Fowler, P. W., Ingram, J. C., Carpenter, E. P., Sansom, M. S., McPherson, M. J., Baldwin, S. A. & Iwata, S. (2011). EMBO J. 30, 417–426.  CrossRef CAS PubMed Google Scholar
First citationNishizawa, T., Kita, S., Maturana, A. D., Furuya, N., Hirata, K., Kasuya, G., Ogasawara, S., Dohmae, N., Iwamoto, T., Ishitani, R. & Nureki, O. (2013). Science, 341, 168–172.  Web of Science CrossRef CAS PubMed Google Scholar
First citationParker, J. L., Li, C., Brinth, A., Wang, Z., Vogeley, L., Solcan, N., Ledderboge-Vucinic, G., Swanson, J. M. J., Caffrey, M., Voth, G. A. & Newstead, S. (2017). Proc. Natl Acad. Sci. USA, 114, 13182–13187.  CrossRef CAS PubMed Google Scholar
First citationPettersen, E. F., Goddard, T. D., Huang, C. C., Couch, G. S., Greenblatt, D. M., Meng, E. C. & Ferrin, T. E. (2004). J. Comput. Chem. 25, 1605–1612.  Web of Science CrossRef PubMed CAS Google Scholar
First citationPotterton, L., Agirre, J., Ballard, C., Cowtan, K., Dodson, E., Evans, P. R., Jenkins, H. T., Keegan, R., Krissinel, E., Stevenson, K., Lebedev, A., McNicholas, S. J., Nicholls, R. A., Noble, M., Pannu, N. S., Roth, C., Sheldrick, G., Skubak, P., Turkenburg, J., Uski, V., von Delft, F., Waterman, D., Wilson, K., Winn, M. & Wojdyr, M. (2018). Acta Cryst. D74, 68–84.  Web of Science CrossRef IUCr Journals Google Scholar
First citationPunjani, A., Rubinstein, J. L., Fleet, D. J. & Brubaker, M. A. (2017). Nature Methods, 14, 290–296.  Web of Science CrossRef CAS PubMed Google Scholar
First citationReyes, N., Ginter, C. & Boudker, O. (2009). Nature (London), 462, 880–885.  Web of Science CrossRef PubMed CAS Google Scholar
First citationScheres, S. H. W. & Chen, S. (2012). Nature Methods, 9, 853–854.  Web of Science CrossRef CAS PubMed Google Scholar
First citationTabata, S., Kitago, Y., Fujii, Y., Mihara, E., Tamura-Kawakami, K., Norioka, N., Takahashi, K., Kaneko, M. K., Kato, Y. & Takagi, J. (2018). Protein Expr. Purif. 147, 94–99.  CrossRef CAS PubMed Google Scholar
First citationTan, Y. Z., Baldwin, P. R., Davis, J. H., Williamson, J. R., Potter, C. S., Carragher, B. & Lyumkis, D. (2017). Nature Methods, 14, 793–796.  Web of Science CrossRef CAS PubMed Google Scholar
First citationYamashita, K., Hirata, K. & Yamamoto, M. (2018). Acta Cryst. D74, 441–449.  Web of Science CrossRef IUCr Journals Google Scholar
First citationYernool, D., Boudker, O., Jin, Y. & Gouaux, E. (2004). Nature (London), 431, 811–818.  CrossRef PubMed CAS Google Scholar
First citationZhang, K. (2016). J. Struct. Biol. 193, 1–12.  Web of Science CrossRef CAS PubMed Google Scholar
First citationZhao, Y., Mao, G., Liu, M., Zhang, L., Wang, X. & Zhang, X. C. (2014). Structure, 22, 1152–1160.  CrossRef CAS PubMed Google Scholar
First citationZheng, S. Q., Palovcak, E., Armache, J., Verba, K. A., Cheng, Y. & Agard, D. A. (2017). Nature Methods, 14, 331–332.  CrossRef CAS PubMed Google Scholar
First citationZivanov, J., Nakane, T., Forsberg, B. O., Kimanius, D., Hagen, W. J. H., Lindahl, E. & Scheres, S. H. W. (2018). Elife, 7, e42166.  CrossRef PubMed Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoSTRUCTURAL BIOLOGY
COMMUNICATIONS
ISSN: 2053-230X
Follow Acta Cryst. F
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds