Skip to main content
Biophysical Journal logoLink to Biophysical Journal
. 2014 Jan 21;106(2):390–398. doi: 10.1016/j.bpj.2013.12.005

Active-Site Structure of the Thermophilic Foc-Subunit Ring in Membranes Elucidated by Solid-State NMR

Su-Jin Kang , Yasuto Todokoro , Ikuko Yumen , Bo Shen , Iku Iwasaki , Toshiharu Suzuki §,, Atsushi Miyagi , Masasuke Yoshida §,, Toshimichi Fujiwara , Hideo Akutsu †,‡,
PMCID: PMC3907233  PMID: 24461014

Abstract

FoF1-ATP synthase uses the electrochemical potential across membranes or ATP hydrolysis to rotate the Foc-subunit ring. To elucidate the underlying mechanism, we carried out a structural analysis focused on the active site of the thermophilic c-subunit (TFoc) ring in membranes with a solid-state NMR method developed for this purpose. We used stereo-array isotope labeling (SAIL) with a cell-free system to highlight the target. TFoc oligomers were purified using a virtual ring His tag. The membrane-reconstituted TFoc oligomer was confirmed to be a ring indistinguishable from that expressed in E. coli on the basis of the H+-translocation activity and high-speed atomic force microscopic images. For the analysis of the active site, 2D 13C-13C correlation spectra of TFoc rings labeled with SAIL-Glu and -Asn were recorded. Complete signal assignment could be performed with the aid of the Cαi+1-Cαi correlation spectrum of specifically 13C,15N-labeled TFoc rings. The Cδ chemical shift of Glu-56, which is essential for H+ translocation, and related crosspeaks revealed that its carboxyl group is protonated in the membrane, forming the H+-locked conformation with Asn-23. The chemical shift of Asp-61 Cγ of the E. coli c ring indicated an involvement of a water molecule in the H+ locking, in contrast to the involvement of Asn-23 in the TFoc ring, suggesting two different means of proton storage in the c rings.

Introduction

H+-driven FoF1-ATP synthase plays a major role in energy production in most organisms. It consists of a water-soluble F1 part and a membrane-integrated Fo part (Fig. 1 A). It converts the electrochemical potential generated by the H+ gradient across membranes into the rotation of the c-subunit ring (red in the figure) in Fo and then into that of the γ subunit in F1, or vice versa (1). The c ring comprises 8–15 c subunits depending on the biological species (2–7). The mechanism underlying the energy conversion at Fo is one of the major unresolved issues regarding FoF1-ATP synthase. The proton translocation across Fo is the key process in the energy conversion and consists of five steps, namely, H+ transfer through the first H+ channel, H+ transfer between the a and c subunits, the rotation of the c ring, H+ transfer between the c and a subunits, and H+ transfer through the second H+ channel. The conserved acidic amino acid residue in the c subunit is known to be essential for this process. Therefore, the active-site structure of the c ring, including the conserved acidic residue, plays an important role in the second and fourth steps. The crystal structures of the c rings of H+-driven FoF1-ATP synthases in detergent micelles have been reported for chloroplasts (8), a cyanobacterium (9), Bacillus pseudofirmus (10), and yeast mitochondria (11). Those of F-type (5) and V-type (12,13) Na+-driven ATPases from bacteria have also been reported. In the case of H+-driven FoF1, the structure of the active site, including the conserved acidic amino acid, was found to be different in the abovementioned crystals. This may be due to the different detergent conditions. Therefore, investigation of a membrane-embedded c ring is required to clarify the structure and function of the active site.

Figure 1.

Figure 1

FoF1-ATP synthase and characterization of the TFoc-subunit oligomers. (A) A model of bacterial FoF1-ATP synthase. (B) Blue native PAGE images of the molecular-weight markers and TFoc oligomers after anion-exchange column chromatography. The sample solution included 0.04% Coomassie G-250, 0.15% DM, and 0.02% DDM. (C) ATP-driven H+-translocating activity of TFoF1 reconstituted from the isolated TFoc oligomer, TFoab2, and TF1. H+-translocating activity of proteoliposomes of TFoF1 was analyzed at 42°C by monitoring ACMA fluorescence at 480 nm and excitation at 410 nm. (a) c-oligomer (cell-free) + ab2, (b) c oligomer (cell-free) + ab2 + F1, and (c) c ring (TFoF1) + ab2 + F1. Here, (cell-free) and (TFoF1) indicate the sources of the c subunits. % ACMA represents the level of quenching of ACMA fluorescence.

We previously reported a solution structure of the c subunit of thermophilic Bacillus PS3 H+-driven FoF1-ATP synthase (TFoF1) (14). However, the structure of its c-subunit ring (TFo c ring) has not yet been determined. The number of subunits in the ring was determined to be 10 using a combination of gene engineering and a biochemical method (4). The amino acid sequence of TFo c in one-letter codes is as follows:

  • MHLGV LAAAI10 AVGLG ALGAG20 IGNGL IVSRT30 IEGIA RQPEL40 RPVLQ TTMFI50

  • GVALV EALPI60 IGVVF SFIYL70 GR

Here, E56 is the conserved glutamate. The solution structure comprised two α helices folded in a hairpin style, as found in every crystal c-ring structure. The conserved Glu is thought to accept/release a proton from/to a proton channel upon interaction with the a subunit. The purpose of this work is to elucidate the structure of the active site of the TFo c ring in lipid membranes using solid-state NMR (ssNMR) in combination with the stereo-array isotope specific labeling. The TFoc ring is suitable for such extensive work because of its stability.

The structural analysis of membrane proteins by ssNMR has been extensively carried out (15–18). However, it is still challenging. We performed ssNMR analysis of 13C, 15N uniformly and specifically labeled Foc subunit rings from Escherichia coli (19,20). Although site-specific isotope labeling by chemical synthesis provided well-defined information, chemical synthesis of multiple samples is a risky and tedious task. In contrast, the fully labeled sample suffered from severe signal overlapping. Therefore, we developed a more efficient, specific labeling method in this work. Cell-free expression can provide a variety of methods for isotope labeling and reconstitution of FoF1. We used stereo-array isotope-labeled amino acids (SAIL-AA) (21) and a wheat-germ extract (WGE) system for specific labeling of TFoc rings. The most critical point in cell-free production is the formation of active rings from the synthesized c-subunit monomers. We examined this by referring to the c rings prepared from TFoF1-ATP synthase. Since the active-site structures in crystals in micelles were different from one another, we focused our attention on this difference. We successfully prepared specifically isotope-labeled TFoc rings and obtained structural information about the active site in lipid membranes. Our results provide insights into the chemical mechanism underlying the proton locking in Foc rings and H+ transfer at the interface of the c and a subunits.

Materials and Methods

The experimental details are presented in the Supporting Material.

Synthesis of TFoc in a WGE system

A mutated protein, S2H-TFoc, was synthesized from messenger RNA (mRNA) in a WGE HPOWG translation mixture at 26°C for ∼24 hr with a supply of amino acids from the dialysis buffer according to a previously described method (22). Because only S2H-TFoc was used in this work, the term “TFoc” will be employed instead of S2H-TFoc hereafter. For isotope labeling of Ala, Gly, Val, Glu, and Asn, 13C,15N-labeled ones were used in the amino acid mixture. SAIL-Asn and -Glu were used at 0.12 mM with the other amino acids deuterated. After the reaction, the collected product was solubilized and applied to Ni-NTA chromatography (23). The column was eluted with an imidazole gradient (20–200 mM) in 100 mM KCl, 0.5% sodium deoxycholate (DOC), 0.15% decyl maltoside (DM), and 20 mM HEPES-KOH buffer (pH 7.5), and the TFoc fraction was collected. This fraction was used for reconstitution into membranes in some cases. Otherwise, this fraction was further applied to an anion exchange HiTrap Q (HQ) column. The HQ column was eluted with a NaCl gradient (0–50 mM) in 0.2 mM EDTA, 10 mM Tris-HCl buffer (pH 7.5). The purity of TFoc was confirmed by Tricine-SDS-PAGE and western blotting.

Reconstitution of TFoc oligomers into lipid bilayers

We solubilized 1,2-diperdeuteriomyristoyl-sn-glycero-3-phosphocholine (DMPC-d54, 98% 2H, 95% chemical purity) in a 0.5% DM solution. This solution was slowly titrated into the protein solution (protein concentration: 0.5 mg/mL; lipid/protein (molar ratio) = 24) under mixing at room temperature. The detergent was removed with Bio-Beads SM-2 (20–50 mesh) followed by dialysis against 100 mM KCl, 0.4 mM NaN3, 20 mM HEPES-KOH/Tris-HCl buffer (pH 7.5), and then against 0.4 mM NaN3, 10 mM HEPES-KOH/Tris-HCl buffer (pH 7.5). The collected precipitate was subjected to freeze-and-thaw treatment 10 times. In the case of SAIL-AA-labeled samples, the buffer was replaced with 0.4 mM NaN3, 10 mM deuterated Tris-2HCl buffer (in 2H2O, p2H 7.5).

Activity analysis and high-speed atomic force microscopy imaging

Active TFoF1 was reconstituted into liposomes. Then, the proton translocation activity was measured according to a previously described method (23,24). A laboratory-built, high-speed (HS) atomic force microscope was used (25). Atomic force microscopy (AFM) images were obtained in the tapping mode. A droplet of TFoc/DMPC-d54 liposome solution was deposited on a freshly cleaved mica surface. After incubation for 5 min, the surface was washed with a solution of 10 mM MgCl2, 50 mM KCl, 20 mM Tris-HCl (pH 7.5). The liposomes retained on the surface were subjected to HS-AFM in the same buffer.

ssNMR measurements

NMR measurements were performed with Varian Infinity-plus 600 and 700 spectrometers operating at 14.09 and 16.44 T, respectively. Broadband double- and triple-resonance magic-angle-spinning (MAS) probes for 3.2 and 2.5 mmϕ rotors were used. The rotor size and MAS rate used for measurements of two-dimensional (2D) spectra were 3.2 mm and 12.5 kHz unless otherwise specified. The amount of protein used was ∼3 mg/3.2 mmϕ rotor. The probe temperature was set at 233 K. The sample temperature was approximately −25°C, because the water signal disappeared in the region of the set temperature from −15 to −20°C due to freezing. 2D dipole-assisted rotational resonance (DARR) (26) and 2D Cαi+1-(C′Cα)i correlation (27) spectra under MAS were recorded with a 3 s repetition time. The 13C chemical shift (CS) was referenced to 2,2-dimethylsilapentane-1-sulfonic acid (DSS) by using the methine carbon signal of adamantane at 40.5 ppm relative to DSS.

Results

Cell-free production of TFoc with the WGE system and isolation of TFoc oligomers

In the absence of a detergent and liposomes, a 1.5–2 mg TFoc/mL reaction mixture was obtained as precipitate. However, the precipitate could not be efficiently solubilized by 2% DOC. To improve the recovery, we examined the effect of an additive such as detergent micelles (Tween 20, NP-40, Triton-X, sucrose monolaurate, or dodecyl maltoside (DDM)) or soybean phosphatidylcholine (sPC) liposomes on expression. The best recovery was achieved in the presence of sPC liposomes. Thus, we decided to synthesize TFoc with WGE in the presence of 0.44% sPC liposomes. We also used 2% DOC to promote ring formation in the solubilization step because it has been used to isolate TFoc rings from E. coli membranes (23). To isolate the TFoc rings, we used a virtual His tag. Although TFoc monomer has only one His at the second position, a decamer ring, for example, should have a noncovalent ring His10 tag. The tricine-SDS-PAGE images in Fig. S1 A indicate the feasibility of this His tag in Ni-NTA chromatography. The obtained TFoc oligomer amounted to a 0.8–1.6 mg/ml reaction mixture. Then, this fraction was applied to an HQ column. The Tricine-SDS-PAGE image at each purification step is presented in Fig. S1 A. Blue native PAGE revealed that the obtained oligomers are homogeneous (Fig. 1 B). To determine whether the isolated oligomers are functional TFoc decamers, we measured the H+-translocation activity of FoF1 reconstituted from the isolated TFoc oligomers, the TFoab2 complexes, and TF1 in sPC liposomes (23). The results are presented in Fig. 1 C. An active FoF1 will translocate H+ from the outside to the inside of the liposome at the expense of ATP added to the outside. The acidification of the inside quenches the fluorescence of 9-amino-6-chloro-2-methoxyacridine (ACMA). The quench rate of the reconstituted TFoF1 including cell-free c oligomers was 69.5%, whereas that of reconstituted TFoF1 including the intact TFoc ring isolated from TFoF1 expressed in E. coli membranes was 71%, i.e., the activity of the former was almost the same as that of the latter. The analysis of the activity of mutant TFoF1 incorporating genetically fused TFoc oligomer (from 2-mer to 14-mer) revealed that only TFoF1 incorporating the decamer ring was active (4). Therefore, we can conclude that the TFoc oligomers obtained from the cell-free system are decamer rings as in the case of the intact ones.

Reconstitution of TFoc oligomers into lipid membranes and HS-AFM imaging

Purified TFoc was reconstituted into DMPC-d54 membranes. The detergent was removed by a combination of Bio-Beads treatment and dialysis as described above. The obtained membrane preparation was characterized by sucrose-density gradient (10–44%) ultracentrifugation as in the case of the EFoc ring (19). The membrane fraction gave a sharp single band that was different from the band for pure DMPC liposomes. To elucidate the macroscopic architecture of the TFoc oligomers in DMPC-d54 membranes, we obtained successive HS-AFM images of the preparation on the mica surface in the presence of buffer at room temperature. As a reference, we observed images of the intact TFoc rings in DMPC-d54 membranes. They were isolated from TFoF1 expressed in E. coli membranes and well characterized in our previous study (23). The images are presented in Fig. 2 A. A torus-shaped substance with a pore at the center could be clearly observed. Since the DMPC-d54 membranes are attached to the mica surface, the TFoc rings should be oriented perpendicular to it. However, it was difficult to discriminate the two different bottom ends (the loop and termini sides) of the TFoc cylinder. The ring diameter was measured using the top position of well-defined rings (see Supporting Material for Materials and Methods). The diameter of the TFoc rings expressed in E. coli was 3.8 ± 0.2 nm (average of 35 images). In Fig. 2 B, we can directly confirm the ring structure of the oligomeric TFoc prepared with the WGE cell-free system. Its diameter was 3.9 ± 0.2 nm (average of 40). Therefore, the oligomeric TFoc is indistinguishable from the intact TFoc decamer ring in terms of the macroscopic structure under physiological conditions.

Figure 2.

Figure 2

HS-AFM images of the TFoc-subunit oligomers in DMPC-d54 membranes. The images were obtained in 10 mM MgCl2, 50 mM KCl, and 20 mM Tris-HCl buffer (pH 7.5) at room temperature. (A) TFoc rings isolated from TFoF1 expressed in E. coli membranes. The imaging rates were 1.72 and 0.95 frames per second (f/s), and the scan ranges were 100 × 100 and 50 × 50 nm on the left and right, respectively. (B) TFoc oligomers isolated from the WGE reaction mixture. The rates were 1.03, 1.63, 1.90, and 1.42 f/s on the left, top right, and left and right of the bottom right, respectively. The scan ranges were 200 × 200 and 50 × 50 nm on the left and right, respectively. (C) Successive images of a TFoc oligomer from the cell-free system disrupted by a cantilever (in the blue circle). The time intervals were 0.6, 0.6, 10.8, and 6.4 s, from a to e, respectively. The imaging rate was 2.95 f/s, except for a (1.63 f/s). The scan range was 20 × 20 nm. The length of the curvature of interest in each shot and its difference are presented.

The advantage of HS-AFM is that it allows one to not only obtain images of a c ring under physiological conditions but also follow a change in the same ring as a function of time. As can be seen in Fig. 2 C, a nonring structure was generated from time to time during scanning. Its curvature changed as a function of the scanning time. This could be ascribed to the degradation caused by the cantilever. The change in the curvature length presented in Fig. 2 C can be summarized as 1.2 × N nm (N, integer), suggesting that regular units are removed from the ring by the cantilever. Although there is no direct evidence, the c subunit is a possible candidate for the regular unit. If this is the case, the circumference of the ring should be described as 1.2 × Nring. Here, Nring is the subunit number in the ring. Since the observed circumference is 12.2 nm, 12.2 ≅ 1.2 × Nring. This leads to the most possible Nring = 10. The most possible number is consistent with the TFoc decamer ring, strongly suggesting that the removed regular unit is a c subunit.

ssNMR measurements

Taking advantage of the WGE cell-free system, we prepared specifically labeled TFoc rings reconstituted in DMPC-d54 membranes to analyze its active site. We focused on Glu-56 (E56) and Asn-23 (N23) because we had previously suggested that the side chain of E56 in the C-terminal helix would interact with that of N23 in the N-terminal helix in the absence of the a subunit in the membrane (14). Because of the low yield and low impurity, the preparation obtained from the Ni-NTA chromatography was used for NMR analysis unless otherwise specified. To obtain information on the TFoc structure in general, and site-specific assignment of E56 and N23, Ala, Gly, Val, Glu, and Asn in TFoc were specifically 13C- and 15N-labeled ([13C,15N]AGVEN-TFoc). The other amino acid residues were nonlabeled ones. Because Ala (nine residues), Gly (11 residues), and Val (eight residues) are abundant, they can provide information about the general structure. In contrast, there are only one Asn and three Glus that will provide specific information about the active site. A 2D DARR spectrum of [13C,15N]AGVEN-TFoc under MAS with a mixing time of 15 ms is presented in Fig. 3 A. The spectrum in the aliphatic region is relatively simple and can be readily assigned in terms of amino acids. Because N23 is unique, the assignment is clear on the basis of its spin connectivity (solid lines). The crosspeaks of Ala and Asn can be used to check the intactness of the TFoc-ring structure obtained from the cell-free system, because they were also isolated in the DARR spectrum of the uniformly 13C-labeled TFoc ring obtained from TFoF1 (23). The CSs of N23 Cα, Cβ, and Cγ were identical for these two spectra. The overlapped crosspeaks of Ala Cβ/Cα of the uniformly labeled TFoc ring obtained from TFoF1 are presented in the inset of the figure. The peculiar pattern is identical to that of this spectrum. Although the crosspeaks of Val Cα/Cβ of the uniformly labeled TFoc ring partially overlap those of Pro, they are also similar to those in this spectra. Since most of Ala, Val, and Asn are distributed in the transmembrane region, it is strongly suggested that the ring structure composed of the TFoc monomers produced with the cell-free system is indistinguishable from that of the intact TFoc ring also at atomic resolution.

Figure 3.

Figure 3

2D 13C-13C correlation DARR (A) and 2D Cαi+1-Cαi correlation (B) spectra of [13C, 15N] AGVEN TFoc-rings in DMPC-d54 membranes at 14.09 T. (A) Mixing time = 15 ms. The amino-acid-specific assignment is presented with a one-letter code. The data size was 1024(d1) × 408(d2) for 60 × 60 kHz spectral widths, and n = 232 scans. The inset is the Ala Cβ/Cα region of the DARR spectrum in Yumen et al. (23). (B) Sequential walking (solid and broken lines) and unambiguous assignment of the crosspeaks are presented; assignment of overlapping peaks is in parentheses. The data size was 1024(d1) × 100(d2) for 60 × 50 kHz spectral widths, and n = 840 scans. A squared sine bell with −80° shift and an exponential with 50 Hz broadening factor and linear prediction were used as window functions in Fourier transformation for the d1 and d2 axes, respectively, in A, and the same window functions with 150 Hz broadening factor were used in B.

To obtain sequential information, we recorded a 2D Cαi+1-(C′Cα)i correlation spectrum under MAS of the same sample (Fig. 3 B). Here, the magnetization was transferred from 13Cαi+1 to 13Cα through 15Ni+1 and 13C = Oi. Now, well-resolved crosspeaks could be obtained because of the uniqueness of successive pairs (for example, A11V12) in the sequence. Sequential walking could be performed, for example, from A11V12 to V12G13 through broken lines. The horizontal line crossed the diagonal line at the CS of V12 at the Ci axis. A vertical line starting from this point found the crosspeak of V12G13. There were two crosspeaks, which might be ascribed to V55E56. However, we could identify the correct one because of the sequential connectivity of V55E56 and E56A57 (indicated by solid lines). The other crosspeak was assigned to V52A53, which was the only possible pair in the region. Thus, the Cα CSs of A11, V12, G13, E32, G33, G51, V52, A53, V55, E56, A57, V63, and V64 were obtained unequivocally. From this result, the crosspeak of Glu observed in both ω1/ω2 and ω2/ω1 (aliphatic areas in the DARR spectrum) turned out to be due to E56 Cα/Cβ. The obtained CSs are summarized in Table S1. These CSs and the CS confinement based on the Cα/Cβ crosspeaks in the DARR spectrum suggest that the backbone structure represented by these residues is composed of α-helical conformations as in the cases of the intact TFoc ring (23) and the solution structure (28).

Then, we focused our attention on E56 and N23. To obtain simple and well-resolved spectra with analyzable intensities, we labeled TFoc with SAIL-Glu and -Asn using the cell-free system (SAIL-EN TFoc). The stereochemical labeling sites are available in Kohno and Endo (22). We used deuterated amino acid residues, except for Glu and Asn. DARR spectra of the SAIL-EN TFoc rings in DMPC-d54 membranes, with mixing times of 100 and 50 ms, are presented in Fig. 4, A and B, respectively. Those with mixing times of 15, 200, and 400 ms are presented in Fig. S2, A–C, respectively. In Fig. 4 A, the Cα/Cβ crosspeaks of N23 and E56 can be identified on the basis of the assignment made for the [13C,15N]AGVEN-TFoc-ring signals. The resolution in the carbonyl carbon region was better than that in the aliphatic region. Typical spin connectivity for Cα, Cβ, Cγ, Cδ, and C′ of E56 is represented by solid lines. The CSs of E56 were determined from those crosspeaks. For the assignment of the other two Glu residues, an expansion of the upper half of the carbonyl region is presented in the inset of Fig. 4 A. The two Glus are designated as Ea and Eb for convenience. They are indicated by solid and broken arrows, respectively. Because the two Cβ/C′ crosspeaks indicated by the two arrows in the inset are separated from each other, the assignment (except for Eb Cα) could be performed using the connectivity with them (see Supporting Material for Results). The Eb Cα CS was determined from the isolated Cα/C′ crosspeak in Fig. 4 B. The CS differences of (Ea-E56) and (Eb-E56) for Cα were 1.1 and 0.3 ppm, respectively. Because the CS difference between E32 and E56 in Fig. 3 B was larger than 1.0 ppm, Ea and Eb should be assigned to E32 and E39, respectively. The assignment was also confirmed with the connectivities in the spectra at 15, 200, and 400 ms mixing times (Fig. S2). The assigned CSs are summarized in Table S2. CSs obtained from SAIL signals were corrected using a calibration table kindly provided by Prof. Masatsune Kainosho.

Figure 4.

Figure 4

(A and B) 2D 13C-13C correlation DARR spectra of SAIL-EN TFoc rings in DMPC-d54 membranes at 16.44 T with mixing times of 100 (A) and 50 (B) ms. Connectivities for N23 and E56 are presented as broken and solid lines, respectively, and those for E32 and E39 are indicated by solid and broken arrows, respectively. (A) The top half of the carbonyl carbon region is expanded in the inset. The data size was 1024(d1) × 210(d2) for 70 × 70 kHz spectral widths, and n = 80 scans. (B) A 2.5 mmϕ rotor (1.5 mg protein) was used at a 15 kHz spinning rate. The data size was 2048(d1) × 210(d2) for 100 × 100 kHz spectral widths, and n = 280 scans. The window functions were a squared sine bell with −80° shift (d1) and an exponential with 50 Hz broadening factor and linear prediction (d2).

The CS of E56 Cδ is significantly different from those of E32 and E39. It is 174.5 ppm for E56, but 183.1 and 182.6 ppm for E32 and E39, respectively. According to a report by Gu et al. (29), isotropic CSs of E32 and E39 Cδ are in the range for deprotonated carboxyl groups. In contrast, that of E56 Cδ is basically in the range for protonated carboxyl groups. However, the isotropic CS of the deprotonated carboxyl group was reported to go down to 172 ppm in crystals (29). To make the protonation state of E56 convincing, we performed CS tensor analysis of E56 Cδ. There is a strong correlation between the principal CS tensor elements (δ11, δ22, and δ33) and chemical structures in solid state (29,30), namely, δ11 is larger than 250 ppm for the protonated carboxyl groups. Also, δ22 provides information about the strength of a hydrogen bond involving the carboxyl group. To determine the principal CS tensor elements of E56 Cδ, we measured 1D spectra of the SAIL-EN TFoc ring (purified with NTA-Ni and HQ columns) in membranes at spinning rates of 4.0, 4.3, 4.5, and 5.0 kHz at 16.44 T. The spectra at spinning rates of 5.0 and 4.5 kHz are presented in Fig. 5 A. In spite of the low signal/noise ratio, the intensity change could be followed, as indicated by arrows in the figure. The side-band intensity data for the four spinning rates were analyzed with Herzfeld-Berger analysis software (HBA 1.6.12) (31,32). Details of the analysis are described in the Supporting Material for Results. The obtained principal elements of the CS tensor were (264 ± 11, 139 ± 7, 121 ± 10) ppm. Because δ11 is larger than 250 ppm, we can safely conclude that this carboxyl group is protonated. Furthermore, the value of δ22 suggests that the E56 carboxyl group is involved in weak hydrogen bonding (29).

Figure 5.

Figure 5

Carbonyl regions of 1D and 2D 13C-NMR spectra of SAIL-EN TFoc rings in DMPC-d54 membranes at 16.44 T. (A) 1D spectra at spinning rates of 4.5 (bottom) and 5.0 (top) kHz. The E56 Cδ signals are indicated by arrows. A 2.5 mmϕ rotor was used. The data size was 4096 for 70 kHz spectral width, and n = 2000 scans. The exponential window function with a 100 Hz broadening factor was used. (B) 2D proton-driven, spin-diffusion 13C-13C correlation spectrum. The relevant crosspeaks are connected by solid lines. The mixing time was 700 ms. Other parameters are given in Fig. S2D.

To obtain structural information around E56, we carried out a 1H-driven spin-diffusion experiment on the SAIL-EN TFoc rings with a mixing time of 700 ms. The relevant parts of the spectrum are presented in Fig. 5 B and Fig. S2 D, in which the crosspeaks of N23 Cδ with E56 Cβ, Cγ, and Cδ can be seen, revealing that they are close to each other.

Discussion

Strategy for analyzing the active site of the TFoc ring in membranes

To analyze the active site of the TFoc ring in membranes, we used specific isotope labeling and CP/MAS NMR. Because this is a membrane protein, we had to develop a method for expression with the WGE cell-free system, formation and purification of the rings, and their reconstitution into membranes. To obtain well-defined structural information about the active site by ssNMR, we used SAIL-Glu and -Asn. This work revealed that the combination of expression with the WGE cell-free system in the presence of liposomes and purification in the presence of 2% DOC could provide functionally active TFoc-decamer rings. Cotranslational insertion of membrane proteins into liposomes in the WGE cell-free system was also previously reported (33). Although TFocs were artificially produced in the cell-free system, they assembled into decamer rings as in the case of TFoF1 formation in vivo. This was verified by the proton-translocating activity, and warranted by the macroscopic and microscopic structures based on HS-AFM images and NMR spectra, respectively. This fact strongly suggests that the number of c subunits that form a ring in lipid membranes is determined by their primary sequence.

The SAIL-EN TFoc rings in the DMPC membranes provided high-resolution ssNMR spectra and structural information in combination with the [13C,15N]AGVEN-TFoc rings. The sensitivity was better for the SAIL-EN TFoc ring than for the AGVEN-labeled one in spite of similar intrinsic line widths (see Supporting Material for Discussion). The better sensitivity likely is due to the higher efficiency in CP and DARR for the SAIL-EN TFoc rings. The longer T1ρ of protons and carbons would improve CP efficiency, and the longer T2 and T1 of carbons may suppress the decay in the evolution and mixing periods of DARR, respectively (26,34,35). The efficient DARR in the presence of only a limited number of protons revealed that the combination of a single 13C-1H group for heterogeneous broadening and a single 1H in the direct neighbor for homogeneous broadening is the core spin system for generating an effective 13C polarization transfer through DARR. The method developed in this work also can be applied to focused analysis of other membrane proteins in lipid membranes.

Structure and function of the active sites of Foc rings in membranes

The active-site structure of the c subunit involving the conserved acidic amino acid residue is a key factor in elucidating the proton-transfer mechanism at the interface between the a and c subunits in Fo. This is also important for understanding the mechanism of the proton translocation across Fo. Our results provide structural information about the active site of the TFoc ring embedded in membranes. The CS of E56 Cδ showed that its carboxyl group was protonated in membranes. The presence of the crosspeaks between N23 Cγ and E56 Cδ, Cγ, and Cβ in the spin diffusion spectrum revealed that the side chain of E56 points to the inside of the ring and interacts with N23, forming a proton-locked conformation. Furthermore, δ22 of E56 Cδ suggested that the carboxyl group is involved in weak hydrogen bonding. This is also supported by the unique isotropic CS of N23 Cγ (170.6 ppm), which is smaller than the average value found in the BMRB database (176.5 ppm). Actually, the isotropic CS of an amide carbon was reported to be proportional to the length of a hydrogen bond in peptide bonds because of a change in δ22 (36), i.e., the smaller the CS, the longer the hydrogen bond.

Our results can be compared with the active-site structures of the Foc rings reported thus far. Actually, they were all different from one another. The cyanobacterium c15 assumed a proton-locked conformation with the conserved carboxyl group on the C-terminal (outer) helix, forming hydrogen bonds with a glutamine side chain on the N-terminal (inner) helix and others (9). Here, the protonated carboxyl group is locked by these hydrogen bonds. The chloroplast c14 assumed a modified proton-locked conformation, missing the hydrogen bond with the glutamine in spite of its presence in the N-terminal helix (8), and Bacillus pseudofirmus c13 assumed a water-involved ion-locked conformation, with the carboxyl group hydrogen bonding only with a water molecule (10). In this case, there is no polar amino acid residue in the active-site region of the N-terminal helix. The presence of the water molecule is similar to the Na+-coordination structure in Foc11 from Ilyobactor tartaricus (5). Since the acidic side chains pointed into the rings in all three cases, they were specified as closed conformations in general. In contrast to them, the conserved carboxyl group (E59) of the yeast mitochondrial c10 (YFoc10) was located outside of the ring in both the protonated and deprotonated states (11). They are called open conformations. The reason for the open-form formation was attributed to the property of detergent by molecular-dynamics simulation. Actually, the YFoc10 E59 exhibited a closed form in a crystal structure of the YF1c10 complex (37). Our work reveals that the closed form is the correct conformation in lipid membranes. The open form was assumed to be a conformation at the interface between the a and c subunits (11).

Furthermore, classification of the c subunits as an E/D-only type or mixed type on the basis of their primary sequence was proposed to be useful for understanding the active-site structure (10). In the former, the conserved Glu or Asp is the only residue that can be predicted to be involved in ion coordination, whereas the latter has more polar residues. To make the difference clearer, we use the term “E/D-plus type” instead of mixed type in this discussion. It turned out so far that the closed conformation in the E/D-only-type ring (c13) carries a water molecule coordinated to the carboxyl group, whereas those in the E/D-plus-type rings (c14 and c15) do not in the high-resolution crystal structures.

Although the whole structure of TFoc10 is not yet known, our result clearly reveals that its conserved carboxyl group is protonated and assumes a proton-locked conformation in lipid membranes. Our result is not consistent with the modified proton-locked conformation of c14, because of the N23/E56 interaction. TFoc10 is the simplest example of the E/D-plus type, since it has only N23 and E56 as the polar residues in the active-site pocket. The protonated state of E56 in TFoc10 and the weak nature of its hydrogen bonding to N23 strongly suggest the absence of a water molecule in this pocket, as in the cases of c14 and c15 (Fig. 6 A). In contrast, the E. coli Foc-subunit ring (EFoc10) is classified to the E/D-only type, the structure of which is also not yet known. It carries no polar residue other than the conserved D61 in the active-site pocket. We performed an ssNMR analysis of EFoc10 in lipid membranes (20). The CS of D61 Cγ was 179.6 ppm. Because this was determined using a chemically synthesized EFoc with specific 13C labeling at A24 Cβ and E56 Cδ, this value is reliable. This CS is significantly different from the 174.5 ppm of E56 Cδ determined in this work. This fact reveals that the chemical nature of the conserved carboxyl group of EFoc10 is significantly different from that of TFoc10. The CS of D61 Cγ suggests that its carboxyl group is either protonated and involved in a strong hydrogen bonding, or is deprotonated and involved in hydrogen bonding (29). Since there is no polar residue to form hydrogen bonds or to coordinate a cation in the active-site pocket, the only possible hydrogen-bonding partner should be a water molecule, as in the case of c13 (Fig. 6 B). The involvement of water is also consistent with the CSs of propionic acid and butyric acid diluted in water, namely, 180.02 and 179.22 ppm for COOH, respectively, and 184.95 and 184.10 ppm for COO, respectively (38). The shorter side chain of D61 may provide a space for the water in the pocket. The presence of water in the pocket may also explain the surviving activity of EFoc10 on the double mutations A24D/D61G and A24D/D61N (39). Although the essential carboxyl group is located on the inner helix in this case, the hydrogen-bonded water molecule would mediate the proton transfer between D24 and the proton channels, responding to the electrostatic interaction with the conserved Arg in the a subunit. Now, we can conclude that the correlation between the water involvement in proton locking and the E/D-only type also holds for EFoc10. This strongly suggests that there are two means of proton storage in the c rings in lipid membranes, namely, polar-group-involved proton locking and water-involved proton locking, as presented in Fig. 6. The weak hydrogen bond in TFoc10 and the presence of the open form in the protonated state in YFoc10 strongly suggest that the barrier between the closed and open forms in the protonated state is low enough to facilitate the closed/open conversion through thermal fluctuations. The major form will be determined by the hydrophilicity of the environment induced by the interaction with the a subunit or lipids. The deprotonation and protonation probably take place in the open form. In the water-involved type, however, there is a possibility that the carboxyl group leaves the proton in the pocket upon conversion from the closed to the open form under the effect of the positive charge in the a subunit, because the water molecule can stabilize it. The deprotonation in the active site generates an ion pair similar to the Na+-locking type. If this is the case, they may have similar mechanisms for cation transfer. There may be also an effect of F1 on the structure of the active site, which we cannot evaluate at this point.

Figure 6.

Figure 6

Two types of proton locking in c10 rings. Sliced views of the active-site pocket from the cytoplasmic side are presented. (A) Polar-group-involved proton locking (TFoc10). The inset is a side view of the model. (B) Water-involved proton locking (EFoc10).

In summary, we have elucidated the chemical nature of the conserved acidic amino acid residue of the TFoc10 and EFoc10 rotors in lipid membranes using ssNMR in combination with the stereo-array isotope specific labeling. It was revealed that the conserved E56 of the former assumed a proton-locked conformation through an interaction with N23. However, D61 of the latter was involved in the interaction with a water molecule. In the case of TFoc10, the fluctuation of the E56 side chain between the closed and open forms could easily take place because of weak hydrogen bonding with N23.

Acknowledgments

We thank Profs. Y. Goto and S. Aimoto at the Institute for Protein Research, Osaka University (IPR-OU), for their assistance in the activity measurements and cleaning of the Biobeads, respectively. The encouragement of Profs. K. Morikawa (IPR-OU) and T. Ando (Kanazawa University) in the HS-AFM experiments is appreciated. We are also grateful to Prof. M. Kainosho (Nagoya University) for fruitful discussions and providing us with a CS calibration table for SAIL-AAs. Thanks are also due to Prof. B.-J. Lee at Seoul National University for his assistance in the NMR analysis.

This work was partly supported by the Targeted Proteins Research Program (H.A., T.F., and M.Y.) and a WCU grant from the Korean Research Foundation (H.A.). S.K. was a member of an international collaborative research project of IPR-OU from 2011 to 2013.

Footnotes

Yasuto Todokoro’s present address is Faculty of Science, Osaka University, Toyonaka, Japan

Atsushi Miyagi’s present address is INSERM U-1006, S/C Unit.ANi 1090/TAGC, 13288, Marseille, France

Supporting Material

Document S1. Figures S1–S3 and Tables S1–S3, References (40–44), detail of Materials and Methods, PAGE images of the preparations in expression and purification, and relevant NMR data and analysis
mmc1.pdf (530.2KB, pdf)
Document S2. Article plus Supporting Material
mmc2.pdf (1.9MB, pdf)

References

  • 1.Yoshida M., Muneyuki E., Hisabori T. ATP synthase—a marvellous rotary engine of the cell. Nat. Rev. Mol. Cell Biol. 2001;2:669–677. doi: 10.1038/35089509. [DOI] [PubMed] [Google Scholar]
  • 2.Stock D., Leslie A.G., Walker J.E. Molecular architecture of the rotary motor in ATP synthase. Science. 1999;286:1700–1705. doi: 10.1126/science.286.5445.1700. [DOI] [PubMed] [Google Scholar]
  • 3.Seelert H., Poetsch A., Müller D.J. Structural biology. Proton-powered turbine of a plant motor. Nature. 2000;405:418–419. doi: 10.1038/35013148. [DOI] [PubMed] [Google Scholar]
  • 4.Mitome N., Suzuki T., Yoshida M. Thermophilic ATP synthase has a decamer c-ring: indication of noninteger 10:3 H+/ATP ratio and permissive elastic coupling. Proc. Natl. Acad. Sci. USA. 2004;101:12159–12164. doi: 10.1073/pnas.0403545101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Meier T., Polzer P., Dimroth P. Structure of the rotor ring of F-type Na+-ATPase from Ilyobacter tartaricus. Science. 2005;308:659–662. doi: 10.1126/science.1111199. [DOI] [PubMed] [Google Scholar]
  • 6.Matthies D., Preiss L., Meier T. The c13 ring from a thermoalkaliphilic ATP synthase reveals an extended diameter due to a special structural region. J. Mol. Biol. 2009;388:611–618. doi: 10.1016/j.jmb.2009.03.052. [DOI] [PubMed] [Google Scholar]
  • 7.Watt I.N., Montgomery M.G., Walker J.E. Bioenergetic cost of making an adenosine triphosphate molecule in animal mitochondria. Proc. Natl. Acad. Sci. USA. 2010;107:16823–16827. doi: 10.1073/pnas.1011099107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Vollmar M., Schlieper D., Groth G. Structure of the c14 rotor ring of the proton translocating chloroplast ATP synthase. J. Biol. Chem. 2009;284:18228–18235. doi: 10.1074/jbc.M109.006916. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Pogoryelov D., Yildiz O., Meier T. High-resolution structure of the rotor ring of a proton-dependent ATP synthase. Nat. Struct. Mol. Biol. 2009;16:1068–1073. doi: 10.1038/nsmb.1678. [DOI] [PubMed] [Google Scholar]
  • 10.Preiss L., Yildiz O., Meier T. A new type of proton coordination in an F(1)F(o)-ATP synthase rotor ring. PLoS Biol. 2010;8:e1000443. doi: 10.1371/journal.pbio.1000443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Symersky J., Pagadala V., Osowski D., Krah A., Meier T., Faraldo-Gomez J.D., Mueller D.M. Structure of the c10 ring of the yeast mitochondrial ATP synthase in the open conformation. Nat. Struct. Mol. Biol. 2012;19:485–491. doi: 10.1038/nsmb.2284. S481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Murata T., Yamato I., Walker J.E. Structure of the rotor of the V-type Na+-ATPase from Enterococcus hirae. Science. 2005;308:654–659. doi: 10.1126/science.1110064. [DOI] [PubMed] [Google Scholar]
  • 13.Mizutani K., Yamamoto M., Murata T. Structure of the rotor ring modified with N,N’-dicyclohexylcarbodiimide of the Na+-transporting vacuolar ATPase. Proc. Natl. Acad. Sci. USA. 2011;108:13474–13479. doi: 10.1073/pnas.1103287108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Nakano T., Ikegami T., Akutsu H. A new solution structure of ATP synthase subunit c from thermophilic Bacillus PS3, suggesting a local conformational change for H+-translocation. J. Mol. Biol. 2006;358:132–144. doi: 10.1016/j.jmb.2006.01.011. [DOI] [PubMed] [Google Scholar]
  • 15.Cady S.D., Schmidt-Rohr K., Hong M. Structure of the amantadine binding site of influenza M2 proton channels in lipid bilayers. Nature. 2010;463:689–692. doi: 10.1038/nature08722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Renault M., Bos M.P., Baldus M. Solid-state NMR on a large multidomain integral membrane protein: the outer membrane protein assembly factor BamA. J. Am. Chem. Soc. 2011;133:4175–4177. doi: 10.1021/ja109469c. [DOI] [PubMed] [Google Scholar]
  • 17.Judge P.J., Watts A. Recent contributions from solid-state NMR to the understanding of membrane protein structure and function. Curr. Opin. Chem. Biol. 2011;15:690–695. doi: 10.1016/j.cbpa.2011.07.021. [DOI] [PubMed] [Google Scholar]
  • 18.Park S.H., Das B.B., Opella S.J. Structure of the chemokine receptor CXCR1 in phospholipid bilayers. Nature. 2012;491:779–783. doi: 10.1038/nature11580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Kobayashi M., Struts A.V., Akutsu H. Fluid mechanical matching of H+-ATP synthase subunit c-ring with lipid membranes revealed by 2H solid-state NMR. Biophys. J. 2008;94:4339–4347. doi: 10.1529/biophysj.107.123745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Todokoro Y., Kobayashi M., Akutsu H. Structure analysis of membrane-reconstituted subunit c-ring of E. coli H+-ATP synthase by solid-state NMR. J. Biomol. NMR. 2010;48:1–11. doi: 10.1007/s10858-010-9432-x. [DOI] [PubMed] [Google Scholar]
  • 21.Kainosho M., Torizawa T., Güntert P. Optimal isotope labelling for NMR protein structure determinations. Nature. 2006;440:52–57. doi: 10.1038/nature04525. [DOI] [PubMed] [Google Scholar]
  • 22.Kohno T., Endo Y. Production of protein for nuclear magnetic resonance study using the wheat germ cell-free system. Methods Mol. Biol. 2007;375:257–272. doi: 10.1007/978-1-59745-388-2_13. [DOI] [PubMed] [Google Scholar]
  • 23.Yumen I., Iwasaki I., Akutsu H. Purification, characterization and reconstitution into membranes of the oligomeric c-subunit ring of thermophilic FoF1-ATP synthase expressed in Escherichia coli. Protein Expr. Purif. 2012;82:396–401. doi: 10.1016/j.pep.2012.02.005. [DOI] [PubMed] [Google Scholar]
  • 24.Suzuki T., Murakami T., Yoshida M. FoF1-ATPase/synthase is geared to the synthesis mode by conformational rearrangement of epsilon subunit in response to proton motive force and ADP/ATP balance. J. Biol. Chem. 2003;278:46840–46846. doi: 10.1074/jbc.M307165200. [DOI] [PubMed] [Google Scholar]
  • 25.Ando T., Kodera N., Toda A. A high-speed atomic force microscope for studying biological macromolecules. Proc. Natl. Acad. Sci. USA. 2001;98:12468–12472. doi: 10.1073/pnas.211400898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Takegoshi K., Nakamura S., Terao T. 13C-1H dipolar-assisted rotational resonance in magic-angle spinning NMR. Chem. Phys. Lett. 2001;344:631–637. [Google Scholar]
  • 27.Fujiwara T., Todokoro Y., Akutsu H. Signal assignments and chemical-shift structural analysis of uniformly 13C, 15N-labeled peptide, mastoparan-X, by multidimensional solid-state NMR under magic-angle spinning. J. Biomol. NMR. 2004;28:311–325. doi: 10.1023/B:JNMR.0000015377.17021.b0. [DOI] [PubMed] [Google Scholar]
  • 28.Wishart D.S., Sykes B.D., Richards F.M. The chemical shift index: a fast and simple method for the assignment of protein secondary structure through NMR spectroscopy. Biochemistry. 1992;31:1647–1651. doi: 10.1021/bi00121a010. [DOI] [PubMed] [Google Scholar]
  • 29.Gu Z.T., Zambrano R., McDermott A. Hydrogen-bonding of carboxyl groups in solid-state amino-acids and peptides—comparison of carbon chemical shielding, infrared frequencies, and structures. J. Am. Chem. Soc. 1994;116:6368–6372. [Google Scholar]
  • 30.Gu Z., Drueckhammer D.G., McDermott A. Solid state NMR studies of hydrogen bonding in a citrate synthase inhibitor complex. Biochemistry. 1999;38:8022–8031. doi: 10.1021/bi9813680. [DOI] [PubMed] [Google Scholar]
  • 31.Herzfeld J., Berger A.E. Sideband intensities in NMR spectra of samples spinning at the magic angle. J. Chem. Phys. 1980;73:6021–6030. [Google Scholar]
  • 32.Eichele K., Wasylishen R.E. Dalhouse University and Universitat Tubingen; 2010. HBA 1.6.12. [Google Scholar]
  • 33.Nozawa A., Ogasawara T., Endo Y. Production and partial purification of membrane proteins using a liposome-supplemented wheat cell-free translation system. BMC Biotechnol. 2011;11:35. doi: 10.1186/1472-6750-11-35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Takegoshi K., Terao T. 13C nuclear Overhauser polarization nuclear magnetic resonance in rotating solids: Replacement of cross polarization in uniformly 13C labeled molecules with methyl groups. J. Chem. Phys. 2002;117:1700–1707. [Google Scholar]
  • 35.Takegoshi K., Nakamura S., Terao T. 13C-1H dipolar-driven 13C-13C recoupling without 13C rf irradiation in nuclear magnetic resonance of rotating solids. J. Chem. Phys. 2003;118:2325–2341. [Google Scholar]
  • 36.Asakawa N., Kuroki S., Ozaki T. Hydrogen-bonding effect on 13C NMR chemical shifts of L-alanine residue carbonyl carbons of peptides in the solid state. J. Am. Chem. Soc. 1992;114:3261–3265. [Google Scholar]
  • 37.Dautant A., Velours J., Giraud M.-F. Crystal structure of the Mg·ADP-inhibited state of the yeast F1c10-ATP synthase. J. Biol. Chem. 2010;285:29502–29510. doi: 10.1074/jbc.M110.124529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Cistola D.P., Small D.M., Hamilton J.A. Ionization behavior of aqueous short-chain carboxylic acids: a carbon-13 NMR study. J. Lipid Res. 1982;23:795–799. [PubMed] [Google Scholar]
  • 39.Miller M.J., Oldenburg M., Fillingame R.H. The essential carboxyl group in subunit c of the F1Fo ATP synthase can be moved and H+-translocating function retained. Proc. Natl. Acad. Sci. USA. 1990;87:4900–4904. doi: 10.1073/pnas.87.13.4900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Koga H., Misawa S., Shibui T. A wheat embryo cell-free protein synthesis system not requiring an exogenous supply of GTP. Biotechnol. Prog. 2009;25:1322–1327. doi: 10.1002/btpr.230. [DOI] [PubMed] [Google Scholar]
  • 41.Morita E.H., Shimizu M., Kohno T. A novel way of amino acid-specific assignment in 1H-15N HSQC spectra with a wheat germ cell-free protein synthesis system. J. Biomol. NMR. 2004;30:37–45. doi: 10.1023/B:JNMR.0000042956.65678.b8. [DOI] [PubMed] [Google Scholar]
  • 42.Bennett A.E., Rienstra C.M., Griffin R.G. Heteronuclear decoupling in rotating solids. J. Chem. Phys. 1995;103:6951–6958. [Google Scholar]
  • 43.Wu G., Wasylishen R.E., Baccolini G. P-31 chemical shieldings in a fused cis-1,2,3-benzothiadiphosphole—a dipolar NMR study. Can. J. Chem. 1992;70:1229–1235. [Google Scholar]
  • 44.Nakai T., Mcdowell C.A. An analysis of NMR spinning side-band of homonuclear 2-spin systems using Floquet theory. Mol. Phys. 1992;77:569–584. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Document S1. Figures S1–S3 and Tables S1–S3, References (40–44), detail of Materials and Methods, PAGE images of the preparations in expression and purification, and relevant NMR data and analysis
mmc1.pdf (530.2KB, pdf)
Document S2. Article plus Supporting Material
mmc2.pdf (1.9MB, pdf)

Articles from Biophysical Journal are provided here courtesy of The Biophysical Society

RESOURCES