Abstract
Anthrax lethal toxin is used as a model system to study protein translocation. The toxin is composed of a translocase channel, called protective antigen (PA), and an enzyme, called lethal factor (LF). A proton gradient (ΔpH) can drive LF unfolding and translocation through PA channels; however, the mechanism of ΔpH-mediated force generation, substrate unfolding, and establishment of directionality are poorly understood. One recent hypothesis suggests that the ΔpH may act through changes in the protonation state of residues in the substrate. Here we report the charge requirements of LF's amino-terminal binding domain (LFN) using planar lipid bilayer electrophysiology. We found that acidic residues are required in LFN to utilize a proton gradient for translocation. Constructs lacking negative charges in the unstructured presequence of LFN translocate independently of the ΔpH driving force. Acidic residues markedly increase the rate of ΔpH-driven translocation, and the presequence is optimized in its natural acidic residue content for efficient ΔpH-driven unfolding and translocation. We discuss a ΔpH-driven charge state Brownian ratchet mechanism for translocation, where glutamic and aspartic acid residues in the substrate are the “molecular teeth” of the ratchet. Our Brownian ratchet model includes a mechanism for unfolding and a novel role for positive charges, which we propose chaperone negative charges through the PA channel during ΔpH translocation.
Keywords: Bacterial Toxins, Biophysics, Molecular Motors, Protein Folding, Protein Stability, Protein Translocation, Brownian Ratchet, Anthrax Toxin, pH Gradient, Proton Motive Force
Introduction
Transmembrane protein translocation (1–4) and intracellular protein degradation (5) are essential processes that allow the cell to traffic protein, form new organelles, maintain protein quality control, and regulate the cell cycle. Because folded proteins are thermodynamically stable under typical cellular conditions, protein translocation and degradation often require complex, energy-consuming molecular machines to catalyze the required series of unfolding and translocation steps in these pathways. Some of these protein transporters and degradation machines mechanically unfold their protein substrates via an ATP-dependent motor. However, a transmembrane proton gradient (ΔpH) may also be used to generate an appropriate driving force that can unfold (6) and translocate a substrate protein across a membrane (1).
A transmembrane proton gradient forms a proton motive force (PMF)2 comprised of two different types of available free energy: an electrical free energy (ΔGΔψ) dependent on the membrane potential (Δψ) and a chemical potential (ΔμΔpH) dependent on the proton concentration gradient (ΔpH). The former ΔGΔψ can be derived from the charge of the translocating protein, z, and Faraday's constant, F, according to the relation, ΔGΔψ = ΔGΔψ° + zFΔψ (1, 6, 7). The latter ΔμΔpH is expressed as a chemical potential, ΔμΔpH = ΔμΔpH° + 2.3 nRT ΔpH (where n is the number of protons involved in the reaction, and R and T are the gas constant and temperature, respectively) (1, 6). Either of these energies can develop enough force under physiological conditions, on the order of tens of pN (6, 8), to unfold a protein during translocation (6, 9). The molecular mechanism of ΔpH-driven translocation, in particular the enforcement of directionality and substrate unfolding, is not well understood.
Anthrax toxin (10) is a useful model system (4) to study ΔpH-driven protein unfolding (11) and translocation (1, 2, 6, 12). The toxin is a key virulence factor secreted by Bacillus anthracis (the causative agent of anthrax). It is comprised of three proteins: the translocase component, protective antigen (PA, 83 kDa); and two enzyme components, lethal factor (LF, 91 kDa) and edema factor (EF, 89 kDa). These secreted proteins must assemble into holotoxin complexes either on the surface of cells or in the blood serum (13). In each assembly mechanism, PA is initially cleaved by a protease, allowing a small 20-kDa portion to dissociate. The remaining 63-kDa PA fragment can then self-assemble into either a ring-shaped homoheptamer, PA7 (14–16), or homooctamer, PA8 (12, 13). PA oligomers then bind to LF and EF and form toxic complexes. Binding occurs at the top of the prechannel and in a deep amphipathic cleft, termed the α clamp, which also aids in unfolding the substrate (17).
The complex is then endocytosed by the target host cell, and subsequent acidification of the endosome causes the PA oligomer to convert into a transmembrane channel. Moreover, the ΔpH generated by the acidification of the endosome facilitates the unfolding (11) and translocation of LF and EF through PA (1) into the cytosol. ΔpH-driven translocation is aided by two catalytic active sites: the α clamp (17) and the Φ clamp (2). The Φ clamp is located inside the PA channel (1, 2). It is comprised of a narrowly apposed ring of seven or eight Phe427 residues, where one Phe is contributed by each PA monomer in the oligomer (2). The Φ clamp prefers to make interactions with hydrophobic and aromatic groups, and its activity facilitates translocation (2) by helping to unfold the substrate (6). How the Φ clamp stimulates (1) ΔpH-driven translocation is unclear.
Although a ΔpH can drive translocation, the groups that are protonated and/or deprotonated in the substrate and/or channel are unknown. One clue is provided by the known cation selectivity of the PA channel (18). This selectivity is likely due to the ∼100–150 glutamic and aspartic acid residues (depending upon the oligomeric state of PA) known to be located inside the PA channel (2, 19). This cation selectivity is unusual given the acidic isoelectric points of both LF and EF (1). Thus we previously proposed a ΔpH-driven charge state Brownian ratchet translocation mechanism (1), which is based upon the theoretical work of Feynman et al. (20), Oster and co-workers (21), and Astumian (22). In this mechanism, the large available thermal energy of the translocating protein is biased in a useful and directed way by the proton gradient, thereby supporting productive translocation.
Because the channel is cation-selective and, therefore, anion-repulsive, we expect that anionic residues within the portion of LF or EF passing through the channel must be protonated, thus making the translocating chain net cationic. The segment of LF within the channel is in fact net positive, as shown by Δψ-dependent measurements, which report positive z values (6). Once the Brownian thermal energy of the translocating protein causes the substrate to exit to the higher pH cytosolic side of the membrane, deprotonation to a more anionic state is favored. The resulting charge repulsion can then enforce directionality during translocation by preventing backward movement through the anionic channel. This process then repeats in multiple cycles toward completion. In this report, we investigate how ΔpH-driven translocation depends upon charged residues in the substrate.
EXPERIMENTAL PROCEDURES
Constructs and Proteins
Site-directed mutagenesis was performed using a QuikChange procedure. WT PA was expressed and purified as described (2). PA7 oligomers were produced as described (12). His6-LFN (residues 1–263 of LF) and mutants thereof were purified from overexpressing bacteria using standard Ni2+-nitrilotriacetic acid affinity chromatography and Q-Sepharose anion exchange chromatography, when required (GE Healthcare) (1). When indicated, His6 tags, which are amino-terminal, 17-residue leaders containing a hexahistidine sequence, MGSSHHHHHHSSGLVPR, were removed from His6-LFN with bovine α-thrombin treatment (0.5 units/mg of protein) for 30 min at room temperature in 20 mm Tris, 150 mm NaCl, 2 mm CaCl2, and 1 m glucose at pH 8.0. Note that after thrombin cleavage, an additional GSHM sequence was left at the amino terminus of all cleaved LFN constructs derived from the pET15b plasmid.
More complex synthetic LFN constructs were made using a three-step, gene synthesis procedure, according to the following scheme: HM1AGGHGDVGM10HVKEKEKNKD20ENKRKDEERN30KTQEEHLKEI40MKHIVKIEVK50GEEAVKKEAA60EKLLEKVPSD70VLEMYKAIGG80KIYIVD. The underlined pairs of amino acids on either end were encoded by the restriction sites, NdeI and a silent SalI site (Val84 and Asp85), respectively, which were used for cloning. Superscripted numbers indicate the numbering convention of LF residues in PDB ID 1J7N (23). Overlapping oligonucleotides encoding the desired sequences with the amino-terminal replacement were synthesized (Elim Biopharmaceuticals, Inc., Hayward, CA) and amplified by two rounds of PCR. In Round I, 22 nm of nested oligonucleotides with consistent annealing temperatures of ∼55 °C were amplified in a standard PCR. In Round II, 1 μl of the PCR product made in Round I was amplified with the two outermost PCR primers (1 μm each) to make the synthetic double-stranded DNA fragment. These LFN synthetic DNA fragments were ligated via a 5′ NdeI site and 3′ SalI site into the pET15b-LFN(SalI) construct, which contains an in-frame, silent SalI restriction site in LFN at Val84 and Asp85. The synthetic LFN constructs were overexpressed and purified, and their His6 tags were subsequently removed as described above (when required).
Electrophysiology
Planar lipid bilayer currents were recorded using an Axopatch 200B amplifier (Molecular Devices Corp., Sunnyvale, CA) (6, 12). Membrane bilayers were painted onto a 100-μm aperture of a 1-ml polyethersulfone cup with 3% 1,2-diphytanoyl-sn-glycero-3-phosphocholine (Avanti Polar Lipids, Alabaster, AL) in neat n-decane. Cis (side to which the PA oligomer is added) and trans chambers were bathed in the indicated buffers as required. By convention, Δψ ψcis − ψtrans (ψtrans 0 V), and ΔpH pHtrans − pHcis.
Translocation Assays
Translocation experiments were carried out as described (6, 12) generally using a universal pH bilayer buffer system (UBB: 10 mm oxalic acid, 10 mm phosphoric acid, 10 mm MES, 1 mm EDTA). However, for translocations requiring a pH > 7.5, we used an altered UBB (6 mm oxalic acid, 6 mm phosphoric acid, 6 mm MES, 6 mm boric acid, 6 mm TAPS, 1 mm EDTA), which is better at buffering in the 7.5–9 range. We found that these two types of buffers produced consistent translocation results. Generally, an additional 100 mm equivalent of KCl was added to the UBB and maintained symmetrically in most translocation experiments, excepting when translocation kinetics were monitored in the absence of a Δψ. In that case, additional KCl was only added to the cis-side buffer. The pHcis was generally 5.6 unless indicated otherwise.
In a typical experiment, we first inserted PA7 channels into a planar bilayer by adding the prechannel oligomeric form to the cis-side of the membrane under a small Δψ of 20 mV. Insertion was observed as an increase in current. Then WT or a mutant LFN was added also to the cis-side of the membrane, and a decrease in current was observed. Excess LFN in the cis-compartment was removed by perfusion, and the translocation process was initiated by either raising the Δψ or by creating a ΔpH by raising the pH of the trans chamber. When translocation is initiated by raising the pH of the trans chamber, we set the time when 1% of maximum translocation occurs as t = 0 s to control for different mixing times.
An alternative method of creating a proton gradient was also used and provided similar results: the pHcis and pHtrans buffers were preadjusted to their final conditions; after substrate binding to the channel was complete, then the cis buffer was perfused at ∼0 mV at pH 5.6 to maintain the ΔpH, and then the voltage was raised to the final voltage upon the completion of perfusion (after ∼30 s).
Single-channel Recordings
Single-channel LFN docking experiments were performed as described (6, 12). Planar lipid bilayers were bathed in symmetric buffer (100 mm KCl, 1 mm EDTA, 10 mm succinic acid, pH 5.6). PA oligomer was applied directly to membranes at ∼10−16 m. Single channel insertion was observed by a discrete step in current under an applied voltage. Once a single channel had inserted into the membrane, LFN was added to the cis-side of the membrane at 200 pm. Blocking events were recorded at a constant Δψ of 20 mV for various LFN mutants with and without their His6 tag. The data were acquired at 400 Hz using a low pass filter of 200 Hz. For LFN that stably blocked a PA channel, the voltage was reversed to −80 mV after several minutes of conductance block to clear the substrate from the channel; this procedure verified that the channel was still present. Histograms of the current versus time data were fit to a two or three Gaussian function to obtain the relative percentages of time spent in the open, blocked, and partly blocked states.
RESULTS
Chemical Potential Component of PMF Is Sufficient to Drive Translocation
To study translocation, we used planar lipid bilayer electrophysiology (1, 2, 6) and a model substrate, LFN, the amino-terminal, 263-residue, PA-binding domain of LF (24). In this assay, a planar lipid bilayer separated two aqueous chambers (called cis and trans). We first inserted PA channels into a planar bilayer by adding the prechannel heptameric form of PA to the cis-side of the membrane under a Δψ of 20 mV (Δψ ψcis − ψtrans, where ψtrans 0 V). Channel formation, observed as an increase in current, then stabilized, and either wild type (WT) LFN or an LFN mutant (MUT) was added also to the cis-side of the membrane. Subsequently, an exponential decrease in current was observed as the amino-terminal presequence of LFN inserted into the ion-conducting channel and blocked ion flow (25). Excess LFN in the cis compartment was removed by perfusion, and the translocation process was initiated by either changing the Δψ and/or ΔpH. The ensuing current increase that was generally observed resulted from LFN translocation to the trans-side of the membrane, as inferred by control experiments (1, 26). Thus two types of parameters can be obtained from translocation records: the time for half of the protein to translocate (t½, measured in seconds) and the efficiency of translocation, which is equivalent to the fraction of LFN that successfully translocates.
Although a ΔpH can drive substrate translocation in the presence of a small, positive Δψ (1), it has not yet been shown whether a ΔpH alone is sufficient. To test this possibility, we set up the bilayer with a potassium chloride gradient; this procedure allowed current to flow when Δψ is 0 mV. It should be noted that by design, the voltage-clamp amplifier maintained the system at 0 mV even under a ΔpH, which we defined as ΔpH pHtrans − pHcis. At a Δψ of 0 mV and positive ΔpH values, we found that both the rate and efficiency of LFN translocation increased with larger positive ΔpH gradients (Fig. 1, inset). We estimated the activation energy of translocation (ΔG‡) using the following relationship, ΔG‡ = RT ln(t½/c), where c is an arbitrary constant of 1 s. The ΔpH dependence of the ΔG‡ for LFN translocation was biphasic with two different limiting slopes (Fig. 1). This dependence is similar to what has been observed under varying ΔpH conditions with a constant, nonzero Δψ (1, 6). The relationship fits to a two-barrier model,
where the indexed ΔG‡o and n values are the activation energy and number of protons required to cross each barrier, respectively (6). The steeper ΔpH dependence of the ΔG‡ on the left-hand side of the graph (less than 1.2 ΔpH, n1 value of 4.0 (±0.5)) corresponds to a barrier limited by LFN unfolding (6), whereas the shallower ΔpH dependence on the right-hand side (at higher ΔpH values, n2 value of 0.22 (±0.03)) may correspond to a barrier limited by a yet uncharacterized translocation process (6) (Fig. 1). We concluded that the chemical potential component of the PMF was sufficient to drive LFN translocation.
ΔpH-driven Translocation Depends on Charged Residues in LFN
The charge state Brownian ratchet model (1) predicted that negatively charged acidic residues in the substrate are critical to ΔpH-driven translocation. To test this model, we replaced the unstructured amino-terminal presequence of LFN (residues 1–27) with a randomized neutral/uncharged background comprised of Gly, Ser and Thr residues (Fig. 2A). We call this neutral mutant background LFN Syn°. This sequence was still polar and flexible but lacked charged residues. We found that LFN Syn° translocated more slowly than LFN WT under conditions of ΔpH = −0.6 to 1.0, pHcis = 5.6, and Δψ = 60 mV (Fig. 2B). Additionally, the construct possessed a significantly reduced ΔpH dependence (LFN WT: n1 = 1.9 (±0.2); LFN Syn°: n = 0.59 (±0.06); Fig. 2B). (It should be noted that we included the amino-terminal His6 tag in these LFN constructs because of defects in channel blocking for certain constructs, an effect that we explore more fully below. To our knowledge, the His6 tag does not obfuscate our results or interpretations, because we were looking at effects relative to similarly His6-tagged control constructs.) Initially, we predicted that acidic residues alone would be sufficient to restore the defect of LFN Syn° in translocation rate and its weaker ΔpH dependence; however, we found this was not the case. LFN Syn−, which has acidic residues restored at their WT positions, further reduced the rate and efficiency of translocation relative to LFN Syn° and had essentially no ΔpH dependence (n value of 0.0 (±0.1); Fig. 2B). Restoring only basic residues in the LFN Syn+ construct increased the translocation rate relative to LFN Syn° but further reduced the ΔpH dependence (n value of 0.14 (±0.07); Fig. 2B). Only when both the acidic and basic residues are restored in LFN Syn± do we see translocation that is identical to the WT substrate (Fig. 2B). Moreover, because LFN WT and LFN Syn± translocated with identical ΔpH dependences, we can safely assume that the noncharged residues in the WT leader sequence are unimportant to ΔpH-driven translocation, as these have been substituted en masse with Gly, Ser, and Thr residues in LFN Syn±. Thus acidic residues alone in the presequence of LFN are not sufficient for efficient ΔpH-driven translocation; instead, a balance of acidic- and basic-charged residues is required.
To further explore the differences between the Syn constructs, we translocated them under more physiologically relevant conditions (ΔpH = 2, Δψ = 0 mV). Suspecting that difficulties in protonating all of the acidic residues in LFN Syn− account for its poor translocation, we also varied the cis pH to see whether lower pHs would increase its translocation speed relative to the chargeless LFN Syn°. Here, we report that the fraction of substrate translocated after 5 min. This is a better method for comparing substrates that display large differences in both translocation efficiencies and rate. Unsurprisingly, LFN WT and LFN Syn± showed no cis pH dependence (Fig. 2C). LFN Syn°, however, displayed a moderate, linear cis pH dependence, likely because of the increased substrate destabilization at lower pHs. As predicted, LFN Syn− had a much stronger cis pH dependence. It displayed little translocation within 5 min at cis pH 5.6 (although it does eventually reach 15–20% translocation). By cis pH 5.0, LFN Syn° and LFN Syn− translocated similarly to LFN WT and LFN Syn±. When the membrane potential was 0 mV, LFN Syn+ translocated poorly at all cis pH levels. We assume that the comparatively better performance of LFN Syn+ relative to LFN Syn− under higher voltage conditions (Fig. 2B) is an artifact of the supplemental Δψ-driving force. Thus we conclude that LFN Syn− is greatly stimulated by a PMF dominated by a ΔpH with lower cis pH conditions, i.e. below ∼pH 5.2.
Cationic Groups in Presequence Allow Anionic Residues to Penetrate Channel
The translocation defect observed for LFN Syn− at pH 5.6 was surprising and led us to hypothesize that a presequence dominated by acidic groups is unable to traverse the cation-selective PA channel, likely because of electrostatic repulsion, which impeded entry into the channel. To test this possibility, we examined the ability of the His6-tagged and untagged versions of the LFN Syn constructs to bind inside the PA channel. In this assay, when the presequence of LFN bound the PA channel, it blocked ion conductance. In bulk, ensemble channel binding assays, we found that all of the His6-Syn constructs could fully engage and dock into the PA channel, as evidenced by the complete block the of channel conductance (Fig. 3A). Upon removing the His6 tag, we found that LFN Syn° and LFN Syn− are defective in forming a stable interaction with the PA channel. Moreover, when we tested the ability of these LFN Syn constructs to remain stably engaged in a blocked conductance state with the PA channel, we found that LFN Syn° and LFN Syn− rapidly dissociated from the channel with dissociation lifetimes of ∼30 and ∼10 s, respectively (Fig. 3B). All other Syn constructs showed little dissociation, and the His6-tagged versions of LFN Syn° and LFN Syn− maintained a stable complex with the PA channel. We conclude that uncharged and purely anionic leader presequences are unable to stably engage the PA channel; however, the defects in these constructs may be complemented by a 17-residue His6 tag.
We then examined the channel blocking activity of the LFN Syn constructs at symmetrical pH 5.6 in single-channel experiments. The His6-tagged Syn constructs all formed stable, fully blocked complexes with the PA channel, where few opening (or unblocking) events were observed (Fig. 3C). However, for the untagged LFN Syn° and LFN Syn− constructs, we again found noticeable defects in channel blocking activity. The two untagged constructs differed notably in their behavior. Untagged LFN Syn° was able to form transient blockades of the PA channel (which fully block the conductance) for durations of 5–10 s. Untagged LFN Syn− is not able to form lasting blockades and only blocked 29 (±2)% of the channel conductance. These partly blocked flickering events lasted 10–30 s (Fig. 3C). From these single-channel studies, we concluded that the two defective mutant LFN presequences (Syn° and Syn−) possess distinguishable characteristics; the purely anionic version (Syn−) cannot penetrate the PA channel and only produces a partial flickering 30% blocked state, whereas the neutral Syn° presequence can completely penetrate and block the channel, albeit ∼70% less often than WT, Syn+, and Syn±. These results, taken together, imply that anionic residues are effectively repulsed by the channel; however, cationic residues can effectively chaperone anionic residues through the channel.
Acidic Residue Dependence of ΔpH Translocation Extends beyond Presequence
Based upon our initial characterization of the sequence dependence of ΔpH-driven translocation, we determined that acidic residues are responsible for the ΔpH dependence, but counter-balancing positively charged residues can assist passage through the cation-selective PA channel. It should be noted that the requirement of acidic residues on the ΔpH dependence of LFN translocation exists when all the negative charges are mutated to serine (supplemental Fig. S1A) and when the His6 tag is cleaved and the negative residues mutated to their uncharged analogs (i.e. Glu → Gln and Asp → Asn; supplemental Fig. S1B). Here we produced a series of amino-terminal LFN mutants in which successively larger amino-terminal stretches of acidic residues are replaced with serine residues (Fig. 4A). We call these anionic residue replacement constructs LFN des(−)x-y, where x and y designate the region of sequence in which the natural Asp and Glu residues were replaced with Ser. Such extensive mutations in LFN destabilized the substrates to varying degrees, as shown in equilibrium protein folding stability measurements (supplemental Fig. S2A). In fact, our analysis is limited to the first 68 residues of LFN because removing acidic residues beyond this point resulted in too substantial a loss in protein solubility and stability.
To normalize for these differences in protein stability, we translocated each His6-tagged LFN des(−) mutant in the presence (i and ii) or absence (iii) of a ΔpH driving force: (i) lower Δψ and higher ΔpH, ΔpH = 1, pHcis = 5.6, Δψ = 20 mV (supplemental Fig. S2B); (ii) higher Δψ and lower ΔpH, ΔpH = 0.6, pHcis = 5.6, Δψ = 40 mV (supplemental Fig. S2C); and (iii) higher Δψ and no ΔpH, ΔpH = 0, pHcis = 5.6, Δψ = 60 mV (supplemental Fig. S2D).
By comparing the translocation rates of a single mutant under two different conditions, we can calculate a ΔΔG‡ value by ΔΔG‡ = ΔG‡(ΔpH>0) − ΔG‡(ΔpH = 0), which normalizes for stability. We then referenced the ΔΔG‡ values for each mutant to that of LFN WT by determining a ΔΔΔG‡ value, where ΔΔΔG‡ = ΔΔG‡(MUT) − ΔΔG‡(WT). Here, larger ΔΔΔG‡ values indicate a reduced capacity to translocate using a pH gradient compared with LFN WT. These comparisons revealed a clear trend showing that the more acidic residues that are removed from LFN, the less it was able to effectively use a pH gradient for translocation (Fig. 4B). Also, we generally found that condition (i), which contains a higher ΔpH, increases ΔΔΔG‡ more so than the lower ΔpH condition (ii), indicating that the effect was specifically due to an inability to use the pH gradient (Fig. 4B). Based on these results, we conclude that acidic residues are critical to the mechanism of ΔpH-dependent translocation both within the presequence region and well into the folded structure of the substrate protein.
Optimal Positions for Acidic Residues in the Presequence
We then asked which acidic residues in the leader presequence were most critical to ΔpH-driven translocation. Using the LFN des(−)1–46 construct as our background, we introduced single glutamates at various positions and measured the translocation kinetics of these new constructs under the influence of a ΔpH. We found that little or no gain (and perhaps a small decrease in rate) was observed when introducing a single Glu at positions 1–13 (inclusive; Fig. 5B). A small but reproducible ∼2-fold increase in translocation rate (relative to the LFN des(−)1–46 background) was observed when introducing a Glu into positions 15–20 (inclusive; Fig. 5B). The optimal region for the introduction of a Glu residue appeared to occur between residues 23 and 31 (inclusive), where the most optimal site was position 23 (increase of 5-fold; Fig. 5B). More modest increases in the translocation rates were observed at residues 34–43 (the extent of the analysis in this mutant background; Fig. 5B). Similar results were obtained in the LFN des(−)1–32 background, and multisite replacements with as few as four of the eight wild-type charges could fully restore the translocation rate to that of LFN WT (supplemental Fig. S3). We concluded that residues 23–31 in the presequence of LFN are the most optimal positions to place acidic residues for ΔpH-mediated translocation.
Anionic Charge Density in Presequence Is Optimized for ΔpH-driven Translocation
We then determined whether the natural sequence of LFN is optimized in terms of the placement of the acidic residues in the presequence. First, we determined the density of various classifications of amino acids: acidic, basic, acidic plus basic, and hydrophobic. Our density metric was calculated as the number of residues in each classification that are ±4 positions, or one turn of an α helix, from the probed site. We then performed linear regression analysis to determine the degree to which the density of these various types of residue classifications correlated with the relative change in translocation rate resulting from the introduction of Glu residues into the LFN des(−)1–46 background (Fig. 5B). We find that acidic residue density provides the strongest correlation with a significant p value of 0.001 (Fig. 5C). Although the density of total charges, acidic plus basic, is a fairly good predictor (p = 0.003), this correlation likely reflects the contribution of the negative charge density and not the positive charge density, because the latter correlation is not significant (p = 0.3). The degree of hydrophobicity also poorly correlated (p = 0.09) with the most optimal positions for acidic residues in ΔpH-mediated translocation. We conclude that the natural placement of acidic residues in the presequence of LFN is optimized for ΔpH-dependent translocation.
Charge-dense Regions Are Optimally Positioned Proximal to Folded Structure
Because we found that when translocating under a ΔpH driving force that the region of highest charge density in the presequence of LFN (residues ∼15 to 30) is also the optimal region for introducing Glu residues into the LFN des(−)1–46 mutant background (Fig. 5), we hypothesized that the location of this cluster of charged residues is only critical in relation to the folded domain of LFN. That is, if the charged cluster were moved further away from the folded structure of LFN, then the force generated by the applied ΔpH gradient would not be as fully realized on the folded structure of LFN. Thus acidic residues more proximal to the folded region of the substrate would better facilitate proper LFN unfolding. To test this hypothesis, we created the LFN 27Ins° construct, which inserts a series of 16 random Gly, Ser, and Thr residues between positions 27 and 28 of LFN WT (Fig. 6A). Despite possessing all of the acidic residues of LFN WT and only extending the overall length of the substrate by ∼6%, translocation is slowed ∼50-fold under a 1.06 unit ΔpH and no Δψ (Fig. 6B). An identical insertion placed earlier (LFN 0Ins° and LFN 11Ins°) in the amino terminus of the LFN did not appreciably effect ΔpH-dependent translocation. Restoring charges to the inserted region in a pattern matching that of LFN residues 12–27 (LFN 27Ins±) returned translocation speed to that of LFN WT. Interestingly, all constructs with inserts (regardless of position, charge, or background) displayed a reduction in efficiency of translocation: ∼50–65% that of LFN WT.
To test whether the reduction of the translocation rate in the LFN 27Ins° construct is due to an effective stabilization of the folded substrate, we introduced the destabilizing mutation, L145A (6), into the LFN WT and 27Ins° backgrounds (Fig. 6C). Under the same conditions described above, the translocation rate of LFN WT was increased by a factor of 2.3 (±0.7) by the L145A mutation, whereas the rate of 27Ins° L145A increases by 6.4 (±1.9). We posit that 27Ins° greatly increases the height of the unfolding barrier (Fig. 6D). When the L145A mutation is introduced in LFN WT, a second barrier becomes limiting, but in 27Ins° the unfolding barrier remains limiting. Thus, despite reducing the height of the unfolding barrier to the same extent, the L145A mutation has a greater effect in 27Ins°. We conclude that the highly charged region must be present immediately prior to the folded domain for efficient protein unfolding to take place under a ΔpH.
Intermixed Cationic and Anionic Groups in the Presequence of LFN are Optimal for ΔpH Translocation
So far we have reported that basic residues are important for the initial insertion into the channel and acidic residues are most important just before the folded domain of LFN. From these observations, we might predict that translocation would not be inhibited when the presequence of LFN is rearranged so that all the basic residues are at the beginning and all the acidic residues are at the end. We made such a construct (LFN PosNeg) along with its inverse (LFN NegPos) and assayed their abilities to translocate with a proton gradient, ΔpH = 1.06 and Δψ = 0 (Fig. 7A). Both constructs were essentially unable to translocate (Fig. 7B). However, a construct with a randomized amino-terminal presequence (LFN Mix) was able to translocate much more similarly to LFN WT (Fig. 7B). Interestingly, none of the positions in this construct share the charge of their WT counterparts. The effectiveness of the LFN Mix construct suggests that the specific WT arrangement of positive and negative charges is less important than simply maintaining an intermixed arrangement of positively and negatively charged residues. We propose that positively charged residues act locally as ionic chaperones for negatively charged acidic residues.
DISCUSSION
To perform work, molecular machines and protein translocases in the cell require a source of energy, generally in the form of chemical energy (ATP) or potential energy stored as either a PMF or other ionic gradient. These sources of energy are interchangeable as evidenced by the highly homologous (27) bacterial flagellum (28, 29), F1Fo-ATP synthase (30, 31), and the vacuolar VoV1-ATPase membrane proton pump (32–34). In the former two cases, a PMF is used to drive the rotational motion of the flagellum (29) or the cycles of enzyme reactions (35) that produce ATP in the transmembrane F1Fo-ATP synthase (36). In the latter system, ATP hydrolysis powers a rotary proton pump to generate a PMF, which is essentially the ATP synthase in reverse. The PMF is also an important driving force for transmembrane protein translocation in mitochondria (37), chloroplasts (38), and the endosomal compartment (1). Proton gradients also exist across the endoplasmic reticulum and other compartments in the cell (39) and in theory may comprise an important driving force for other types of translocases in the cell.
Active Pushing/Pulling Translocation Models
Protein translocation is basically a series of reactions that convert a source of free energy to mechanical work used to drive protein unfolding and unidirectional transport. This process is involved in transmembrane transport, protein degradation, and chaperone activities. Here, either ATP hydrolysis cycles (5, 40–42) or utilization of the PMF (1) can propel polypeptides across membranes or into hydrolytic compartments for degradation. Often the molecular machine that carries out translocation processes contains loop regions with critical aromatic groups at their tips. Also, additional accessory proteins outside the channel may be utilized to engage the polypeptide during translocation. It is thought that cycles of ATP binding, ATP hydrolysis, and ADP and inorganic phosphate release cause these loops/proteins to move like simple linear actuators, which effectively pushes the translocating polypeptide in a particular direction (40, 42–44). This mechanism is supported in part because these Φ-clamp-type structures are found in various conformations for different nucleotide bound states. For anthrax toxin, the utilization of the proton gradient occurs most optimally with an intact aromatic Φ-clamp structure (1, 2), suggesting that these aromatic internal pore loops may push the translocating polypeptide in a similar manner. At present, this active pushing mechanism of translocation is favored in many systems, but it cannot by itself explain all of the mechanochemical coupled mechanisms.
Brownian Ratchet Models for Translocation
An alternative, but potentially complementary view, of mechanochemical coupling should be considered in addition to the active pushing model. The Brownian ratchet model (20–22, 45, 46) suggests that nonequilibrium fluctuations, namely in the form of chemical reactions or thermal gradients, can bias random thermal motion in a productive and directional way. In Feynman's initial conception (20), stochastic fluctuations may be harnessed for useful work using a temperature gradient. Although a thermal gradient is difficult to achieve across a biological membrane, chemical asymmetries and membrane gradients, such as a ΔpH or those created by ATP hydrolysis, are readily available to drive biological Brownian motors.
Ratchet Model of ΔpH-driven Anthrax Toxin Translocation
Anthrax toxin activity is most optimal when the endosomal compartment is allowed to naturally acidify (47). Acidification produces a transmembrane PMF that stimulates translocation (1, 2, 6). The chemical potential component of the gradient, ΔμΔpH, is sufficient by itself to drive translocation (Fig. 1). How might ΔpH-driven translocation follow a Brownian ratchet mechanism? Brownian fluctuations in the substrate polypeptide chain may be coupled to protonation state changes in the substrate chain or channel (Fig. 8). We infer that acidic residues in the substrates (which have net acidic isoelectric points) are protonated upon entering the PA channel, because the charge dependence, or z value, we observe for translocation is positive. This fact implies that the substrate chain is positively charged at the point of the potential drop in the channel (1, 6). The positive charge dependence also reflects the known fact that the PA channel is cation-selective (18), and a highly anionic polypeptide would be repulsed by the channel as we observe with our LFN Syn− construct (Fig. 2B). A similar effect has been observed when LFN is modified with a sulfate group (48). Therefore, a segment of substrate chain with protonated acidic residues will more readily achieve a positive charge enabling its passage through the PA channel under Brownian motion. It can subsequently release protons into the trans-side of the membrane (as favored in the higher pH condition), thus leading to a build-up of anionic charge in the trans-side substrate. The anionic charge in the trans substrate should effectively impede retrograde efflux (under Brownian motion) back into the anion-repulsive channel and enforce proper directionality. Cycles of this charge state Brownian ratchet mechanism would then translocate the remainder of the polypeptide across the membrane.
Acidic Residues in the Substrate Are the Teeth in the ΔpH-driven Brownian Ratchet
One obvious and testable feature of this charge state Brownian ratchet mechanism is the role of acidic residues in the substrate. A previous study using a synthetic peptide attached to the folded domain of LFN via native chemical ligation showed that chargeless or positive-only presequences are defective in translocation (49). Here we more extensively explored charged residue involvement in ΔpH translocation. By replacing the residues in the presequence en masse with neutral residues and measuring translocation under a variety of ΔpH conditions, we determined the ΔpH dependence of translocation, which is the best indicator of the ability to use a proton gradient as a driving force (Fig. 2). We found that sequences lacking acidic residues had comparatively weak ΔpH dependences (Fig. 2) and the ability to translocate using a ΔpH decreases as more acidic residues are removed (Fig. 4). A more detailed scan, using an anion-less variant, LFN des(−)1–46, showed that the region spanning residues 23–31 received the largest gain in translocation kinetics when a single acidic residue was introduced (Fig. 5B). Furthermore, the natural sequence appears to maximize its inherent acidic residue density in this very region to best take advantage of the ΔpH driving force (Fig. 5C). Thus based on this model, we conclude that acidic residues in the substrate form the molecular teeth in the ΔpH-driven Brownian ratchet mechanism.
Novel Role for Positive Charges in the Charge State Ratchet
We identified one other complicating factor in our analysis when we tested for the role of positive charges in the presequence. Using the synthetic presequence Syn series, we find that a Syn− presequence dominated by acidic residues and lacking basic residues is defective in translocation (Fig. 2B). However, this result does not mean that ΔpH-driven translocation requires only positive charges, because presequences engineered to only contain cationic residues tend to have flat ΔpH dependences (Fig. 2B). Various lines of evidence indicate that acidic residues require neutralization either by direct protonation (Fig. 2C) or by proximal positively charged residues (Figs. 3 and 7). This requirement demonstrates that the channel has robust charge selectivity. We propose this selectivity enforces directionality through electrostatic repulsion, thereby making it an integral feature of the charge state Brownian ratchet translocation mechanism.
Can ΔpH-driven Brownian Ratchets Generate Robust Unfoldase Activity?
Some formulations (21, 44), but not all (50), of the Brownian ratchet mechanism imply that unfolding must occur prior to the engagement of the ratcheting process. In the unfolding-limited conceptualization of the Brownian ratchet model, it is suggested that the channel must wait for the protein to unfold and the channel itself is not a participant in the denaturation. In this view, it is thought that the Brownian ratchet mechanism cannot fully explain the large translocation rate accelerations observed for stable substrates, such as dihydrofolate reductase (51), barnase (52), or cytochrome b2 (53).
This formulation of the Brownian ratchet model requires some updating with more current reports (1, 2, 6, 17). The PA channel, for example, utilizes two principle unfoldase or denaturation sites within the channel, namely the Φ clamp (2, 6) and α clamp (17). The former site is comprised of a ring of seven or eight Phe residues (depending on the oligomerization state of the PA complex), which can clamp onto the amino terminus of LF (2) and drive unfolding (6), presumably by binding to hydrophobic moieties in the substrate in a nonspecific manner (2). The latter α-clamp site is a deep cleft, situated between the twin Ca2+ ion-binding sites on the surface of the PA oligomer; this clamp is capable of binding nonspecifically to a ∼10-residue α helix and short ∼5-residue β strand (17). The initial characterization of the α clamp revealed an interesting capability of the site: upon binding to the surface of the PA oligomer, LFN is partially unfolded. In this unfolding, the first α helix and β strand of LF unfurl and dock into the α-clamp cleft. The α clamp also has broad binding specificity, allowing the site to recognize many types of sequences and making it a general denaturation site on the translocase. The second factor to consider is that ΔpH-driven translocation facilitates unfolding (6). Two different types of barriers have been identified in LFN translocation. The more force-dependent barrier (as determined by the n value, Equation 1) is also limited by protein unfolding (6). Thus substrate unfolding may not be out of the realm of possibility for a Brownian ratchet powered translocase, because the closely spaced denaturation sites in the channel allow for small displacements of sequence (small fluctuations) to be captured and stabilized via interactions with the channel.
How Might Denatured Protein-binding Sites on the Channel Operate in Conjunction with a ΔpH-driven Ratchet?
Brownian thermal fluctuations are a significant source of energy that, if partially harnessed, could serve to denature a protein. For a polypeptide presequence (30 residues in length), Brownian thermal fluctuations provide ∼6 kcal mol−1 of thermal energy (assuming two degrees of freedom per backbone Φ and Ψ angle/residue). Thus, although the α-clamp structure maintains contact with the sequence carboxyl-terminal to the presequence, the amino-terminal end may move through the channel purely motivated by Brownian thermal fluctuations. Successful excursions require that the presequence region is electrostatically compatible with the cation-selective lumen of the channel. We believe that this is ensured either through the lowered pH of the cis-side of the membrane or by adjacent positive charges, which may act to help chaperone the acidic groups across the membrane. Successful excursions are then anchored or captured on the trans-side of the membrane once the acidic groups further deprotonate in the higher pH medium. The high degree of correlation between the best sites to introduce a Glu residue in LFN des(−)1–46 and the naturally high density of negative charges in the presequence of LF (Fig. 5C) suggests that this anchoring step is more favorable with higher acidic residue densities.
Interestingly, we find that the acidic residue-rich sequence is located immediately adjacent to the folded structure. Thus after successful anchoring of the presequence takes place, an entropic tension may be effectively applied to the remaining folded structure. This entropic tension is derived from the fact that a more extended polypeptide chain would have fewer possible conformations, and to relieve this entropic tension, the protein would be driven to unfold. We tested the positional dependence of this acidic residue cluster by creating an artificial gap between this naturally dense region and the folded structure of LFN. When the gap contains neutral residues, the translocation rate decreases 50 fold, but when the gap contains charged groups, the translocation rate is similar to WT (Fig. 6B). Furthermore, the decrease in the translocation rate coincides with an observed stabilization of the substrate (Fig. 6C). Previous studies show that a larger nucleus of structure unfolds during translocation in a large cooperative unfolding step (6); however, this unfolding occurs after a smaller portion of the amino-terminal structure is unfolded and docked into the α-clamp site (17). Therefore, we expect that the α- and Φ-clamp protein denaturation sites in the channel can reduce the overall unfolding barrier by allowing for small incremental unfolding steps to be stabilized. Effective trapping of Brownian fluctuations may force partially unfolded intermediates to then disengage from the channel, leading ultimately to the larger scale cooperative unfolding event (6) that we observe under a ΔpH.
Finally, the charge requirements we identify for the substrate strongly favor a Brownian ratchet model. However, we cannot rule out the possibility that ΔpH-driven transport also involves coordinated, proton-dependent movement of loops within the PA channel, which helps to push the substrate chain during translocation. A combination of the active pushing and Brownian ratchet models may apply to this system because these models are not mutually exclusive. In fact, other translocases already known to push proteins using ATP-driven loop movements may also use a Brownian ratchet to further drive transport.
Supplementary Material
Acknowledgments
We greatly appreciate members of the Krantz lab, S. Marqusee, D. Wemmer, and A. Martin for helpful discussions.
This work was supported, in whole or in part, by National Institutes of Health Grant R01-AI077703.
The on-line version of this article (available at http://www.jbc.org) contains supplemental text and Figs. S1–S3.
- PMF
- proton motive force
- PA
- protective antigen
- LF
- lethal factor
- EF
- edema factor
- LFN
- lethal factor amino-terminal PA-binding domain
- Syn
- synthetic sequence region
- Ins
- synthetic insertion sequence
- des(−)
- sequence region lacking acidic charges
- MES
- 4-morpholineethanesulfonic acid
- TAPS
- 3-{[2-hydroxy-1,1-bis(hydroxymethyl)ethyl]amino}-1-propanesulfonic acid
- UBB
- universal bilayer buffer.
REFERENCES
- 1. Krantz B. A., Finkelstein A., Collier R. J. (2006) J. Mol. Biol. 355, 968–979 [DOI] [PubMed] [Google Scholar]
- 2. Krantz B. A., Melnyk R. A., Zhang S., Juris S. J., Lacy D. B., Wu Z., Finkelstein A., Collier R. J. (2005) Science 309, 777–781 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3. Wickner W., Schekman R. (2005) Science 310, 1452–1456 [DOI] [PubMed] [Google Scholar]
- 4. Thoren K. L., Krantz B. A. (2011) Mol. Microbiol. 80, 588–595 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Sauer R. T., Bolon D. N., Burton B. M., Burton R. E., Flynn J. M., Grant R. A., Hersch G. L., Joshi S. A., Kenniston J. A., Levchenko I., Neher S. B., Oakes E. S., Siddiqui S. M., Wah D. A., Baker T. A. (2004) Cell 119, 9–18 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Thoren K. L., Worden E. J., Yassif J. M., Krantz B. A. (2009) Proc. Natl. Acad. Sci. U.S.A. 106, 21555–21560 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. Huang S., Ratliff K. S., Matouschek A. (2002) Nat. Struct. Biol. 9, 301–307 [DOI] [PubMed] [Google Scholar]
- 8. Shariff K., Ghosal S., Matouschek A. (2004) Biophys. J. 86, 3647–3652 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Brockwell D. J., Paci E., Zinober R. C., Beddard G. S., Olmsted P. D., Smith D. A., Perham R. N., Radford S. E. (2003) Nat. Struct. Biol. 10, 731–737 [DOI] [PubMed] [Google Scholar]
- 10. Young J. A., Collier R. J. (2007) Annu. Rev. Biochem. 76, 243–265 [DOI] [PubMed] [Google Scholar]
- 11. Krantz B. A., Trivedi A. D., Cunningham K., Christensen K. A., Collier R. J. (2004) J. Mol. Biol. 344, 739–756 [DOI] [PubMed] [Google Scholar]
- 12. Kintzer A. F., Thoren K. L., Sterling H. J., Dong K. C., Feld G. K., Tang I. I., Zhang T. T., Williams E. R., Berger J. M., Krantz B. A. (2009) J. Mol. Biol. 392, 614–629 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Kintzer A. F., Sterling H. J., Tang I. I., Abdul-Gader A., Miles A. J., Wallace B. A., Williams E. R., Krantz B. A. (2010) J. Mol. Biol. 399, 741–758 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Lacy D. B., Wigelsworth D. J., Melnyk R. A., Harrison S. C., Collier R. J. (2004) Proc. Natl. Acad. Sci. U.S.A. 101, 13147–13151 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Milne J. C., Furlong D., Hanna P. C., Wall J. S., Collier R. J. (1994) J. Biol. Chem. 269, 20607–20612 [PubMed] [Google Scholar]
- 16. Petosa C., Collier R. J., Klimpel K. R., Leppla S. H., Liddington R. C. (1997) Nature 385, 833–838 [DOI] [PubMed] [Google Scholar]
- 17. Feld G. K., Thoren K. L., Kintzer A. F., Sterling H. J., Tang I. I., Greenberg S. G., Williams E. R., Krantz B. A. (2010) Nat. Struct. Mol. Biol. 17, 1383–1390 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Blaustein R. O., Koehler T. M., Collier R. J., Finkelstein A. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 2209–2213 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Nassi S., Collier R. J., Finkelstein A. (2002) Biochemistry 41, 1445–1450 [DOI] [PubMed] [Google Scholar]
- 20. Feynman R. P., Leighton R. B., Sands M. (1963) The Feynman Lectures on Physics, pp. 46-1–46-9, Addison-Wesley, Reading, MA [Google Scholar]
- 21. Simon S. M., Peskin C. S., Oster G. F. (1992) Proc. Natl. Acad. Sci. U.S.A. 89, 3770–3774 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Astumian R. D. (1997) Science 276, 917–922 [DOI] [PubMed] [Google Scholar]
- 23. Pannifer A. D., Wong T. Y., Schwarzenbacher R., Renatus M., Petosa C., Bienkowska J., Lacy D. B., Collier R. J., Park S., Leppla S. H., Hanna P., Liddington R. C. (2001) Nature 414, 229–233 [DOI] [PubMed] [Google Scholar]
- 24. Lacy D. B., Mourez M., Fouassier A., Collier R. J. (2002) J. Biol. Chem. 277, 3006–3010 [DOI] [PubMed] [Google Scholar]
- 25. Zhang S., Finkelstein A., Collier R. J. (2004) Proc. Natl. Acad. Sci. U.S.A. 101, 16756–16761 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Zhang S., Udho E., Wu Z., Collier R. J., Finkelstein A. (2004) Biophys. J. 87, 3842–3849 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Cross R. L., Müller V. (2004) FEBS Lett. 576, 1–4 [DOI] [PubMed] [Google Scholar]
- 28. Sowa Y., Berry R. M. (2008) Q. Rev. Biophys. 41, 103–132 [DOI] [PubMed] [Google Scholar]
- 29. Berg H. C. (2003) Annu. Rev. Biochem. 72, 19–54 [DOI] [PubMed] [Google Scholar]
- 30. Oster G., Wang H. (1999) Structure 7, R67–R72 [DOI] [PubMed] [Google Scholar]
- 31. Abrahams J. P., Leslie A. G., Lutter R., Walker J. E. (1994) Nature 370, 621–628 [DOI] [PubMed] [Google Scholar]
- 32. Gregorini M., Wang J., Xie X. S., Milligan R. A., Engel A. (2007) J. Struct. Biol. 158, 445–454 [DOI] [PubMed] [Google Scholar]
- 33. Nakanishi-Matsui M., Sekiya M., Nakamoto R. K., Futai M. (2010) Biochim. Biophys. Acta 1797, 1343–1352 [DOI] [PubMed] [Google Scholar]
- 34. Finbow M. E., Harrison M. A. (1997) Biochem. J. 324, 697–712 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Boyer P. D. (1997) Annu. Rev. Biochem. 66, 717–749 [DOI] [PubMed] [Google Scholar]
- 36. Noji H., Yasuda R., Yoshida M., Kinosita K., Jr. (1997) Nature 386, 299–302 [DOI] [PubMed] [Google Scholar]
- 37. Voos W., Martin H., Krimmer T., Pfanner N. (1999) Biochim. Biophys. Acta 1422, 235–254 [DOI] [PubMed] [Google Scholar]
- 38. Sargent F. (2007) Biochem. Soc. Trans. 35, 835–847 [DOI] [PubMed] [Google Scholar]
- 39. Llopis J., McCaffery J. M., Miyawaki A., Farquhar M. G., Tsien R. Y. (1998) Proc. Natl. Acad. Sci. U.S.A. 95, 6803–6808 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Wang J., Song J. J., Franklin M. C., Kamtekar S., Im Y. J., Rho S. H., Seong I. S., Lee C. S., Chung C. H., Eom S. H. (2001) Structure 9, 177–184 [DOI] [PubMed] [Google Scholar]
- 41. Lum R., Niggemann M., Glover J. R. (2008) J. Biol. Chem. 283, 30139–30150 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Martin A., Baker T. A., Sauer R. T. (2008) Nat. Struct. Mol. Biol. 15, 1147–1151 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. DeLaBarre B., Brunger A. T. (2005) J. Mol. Biol. 347, 437–452 [DOI] [PubMed] [Google Scholar]
- 44. Glick B. S. (1995) Cell 80, 11–14 [DOI] [PubMed] [Google Scholar]
- 45. Astumian R. D., Bier M. (1994) Phys. Rev. Lett. 72, 1766–1769 [DOI] [PubMed] [Google Scholar]
- 46. Yamano K., Kuroyanagi-Hasegawa M., Esaki M., Yokota M., Endo T. (2008) J. Biol. Chem. 283, 27325–27332 [DOI] [PubMed] [Google Scholar]
- 47. Friedlander A. M. (1986) J. Biol. Chem. 261, 7123–7126 [PubMed] [Google Scholar]
- 48. Basilio D., Juris S. J., Collier R. J., Finkelstein A. (2009) J. Gen. Physiol. 133, 307–314 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Pentelute B. L., Barker A. P., Janowiak B. E., Kent S. B., Collier R. J. (2010) ACS Chem. Biol. 5, 359–364 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Stuart R. A., Cyr D. M., Craig E. A., Neupert W. (1994) Trends Biochem. Sci. 19, 87–92 [DOI] [PubMed] [Google Scholar]
- 51. Matouschek A., Azem A., Ratliff K., Glick B. S., Schmid K., Schatz G. (1997) EMBO J. 16, 6727–6736 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Huang S., Ratliff K. S., Schwartz M. P., Spenner J. M., Matouschek A. (1999) Nat. Struct. Biol. 6, 1132–1138 [DOI] [PubMed] [Google Scholar]
- 53. Glick B. S., Wachter C., Reid G. A., Schatz G. (1993) Protein Sci. 2, 1901–1917 [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.