2,076
Views
34
CrossRef citations to date
0
Altmetric
Original Article

Structure and function of the bacterial Sec translocon (Review)

, &
Pages 387-394 | Received 09 Feb 2007, Published online: 09 Jul 2009

Abstract

Bacteria and archaea possess a protein complex in the plasma membrane that governs protein secretion and membrane protein insertion. Eukaryotes carry homologues in the endoplasmic reticulum (ER) where they direct the same reaction. A combination of experiments conducted on the systems found in all three domains of life has revealed a great deal about protein translocation. The channel provides a route for proteins to pass through the hydrophobic barrier of the membrane, assisted by various partner proteins which maintain an unfolded state of the substrate, target it to the channel and provide the energy and mechanical drive required for transport. In bacteria, the post-translational reaction utilizes an ATPase that couples the free energy of ATP binding and hydrolysis to move the substrate through the protein pore. This review will draw on genetic, biochemical and structural findings in an account of our current understanding of this mechanism.

Introduction

Proteins required in extra-cytosolic locations rely on specific targeting and transport apparatus for their localization. This process directs protein secretion and membrane protein insertion, and is therefore essential for cellular biogenesis. There exists a ubiquitous membrane protein complex found in the plasma membrane of bacteria and archaea (SecY), and in the ER of eukaryotic cells (Sec61). Two modes of translocation (co- and post-translational) converge at the Sec complex, which is a versatile and dynamic structure capable of conducting large substrates through the membrane without the loss of small molecules.

Some of the proteins synthesized in the cytosol carry N-terminal signal sequences, which are recognized by factors required for targeting Citation[1]. In the event of co-translational translocation, the ribosome-nascent chain complex is targeted to the membrane courtesy of the signal recognition particle (SRP) and its receptor Citation[2], Citation[3]. Subsequent reactions deliver the translating ribosome to the SecY/Sec61 complex and in this state the ribosomal polypeptide exit site is located in close vicinity to the Sec complex Citation[4], enabling protein translocation to be driven by the concomitant chain elongation Citation[5].

Post-translational translocation in eukaryotes and bacteria operate by different mechanisms. In eukaryotes this process relies on BiP, an Hsp70 homologue in the ER lumen which utilises ATP to effectively pull the polypeptide chain in a ratchet-like mechanism Citation[6], Citation[7]. Bacteria have adopted a mechanism that pushes proteins across the channel by employing a motor protein ATPase SecA Citation[8], Citation[9]. In addition, the reaction requires a cytosolic component SecB to aid targeting to the membrane and to prevent premature protein folding and aggregation, to maintain a translocation competent conformation of the pre-protein Citation[8], Citation[10–12].

Elements of the Sec pathway were identified through different genetic screens conducted with Escherichia coli. Conditional-lethal mutations associated with a generalized protein-secretion defect facilitated the identification of the genes secA, secB, secE and secYCitation[13–15]. The second strategy selected for mutations restoring the periplasmic localization of proteins with secretion-defective signal sequences, where prlD, prlG and prlA were identified as allelic forms of secA, secE and secY, respectively Citation[16–18]. A landmark advance illustrated the requirement for ATP in post-translational translocation Citation[19] and subsequent analysis of the gene products identified three membrane protein components SecY, SecE and SecG of the protein channel, and SecA a soluble ATPase Citation[9], Citation[20–22]. Seminal experiments reconstituted the entire translocation reaction in vitro, when a Sec dependent substrate could be transported to the interior of vesicles containing SecYEG in the presence of SecA and ATP Citation[9].

The SecY protein channel

Images recorded by electron cryo-microscopy provided the first pictures of the Sec complex and revealed an oligomeric assembly of an estimated 3–4 complexes with an area of low density formed at the interface Citation[23]. This pore aligned with the polypeptide exit channel of the ribosome Citation[24–26]. The first crystal structure determined to a significantly higher resolution was of the membrane bound E. coli SecYEG complex. Electron microscopy and image reconstruction revealed all 15 of the predicted trans-membrane (TM) α-helices, Citation[27], Citation[28]. The arrangement of the SecYEG protomers seen in the membrane bound state was that of a dimer in a so-called ‘back-to-back’ arrangement.

The breakthrough came with the solution of a high-resolution X-ray structure of a monomeric and closed archaeal SecY complex, resolved in detergent solution () Citation[29]. The most surprising observation was a hydrophilic restriction point in the centre of the monomer reminiscent of a closed channel or pore. This narrow pore is held closed by a short reinserted loop (the ‘plug’) and a girdle of hydrophobic side chains. The more conserved residues in SecY line the channel, which is also the site of several prl mutations Citation[29]. This putative channel is conveniently located next to the signal sequence binding pocket between TM2 & 7 Citation[30] and has also been shown to form cross-links with translocating pre-protein Citation[31]. The periphery of the monomeric complex is too hydrophobic to form part of a protein channel at the interface of an oligomeric complex.

Figure 1.  Ribbon and schematic representation of the detergent solubilized monomeric Methanococcus jannaschii SecYEβ complex viewed from the cytosolic face Citation[29]. Transmembrane helices 1–5 are coloured in light green, helices 6–10 dark green, the ‘plug’ orange, SecE red and Secβ (SecG) is shown in sea green. The equivalent E. coli cysteine cross-link (L106C) is indicated in blue, close to the blue arrow. The green arrow denotes the position of the lateral gate.

Figure 1.  Ribbon and schematic representation of the detergent solubilized monomeric Methanococcus jannaschii SecYEβ complex viewed from the cytosolic face Citation[29]. Transmembrane helices 1–5 are coloured in light green, helices 6–10 dark green, the ‘plug’ orange, SecE red and Secβ (SecG) is shown in sea green. The equivalent E. coli cysteine cross-link (L106C) is indicated in blue, close to the blue arrow. The green arrow denotes the position of the lateral gate.

The structure and oligomeric state of the active channel

The channel pore is formed between two distinct halves of the SecY protein (). SecE has been proposed to form a clamp around these domains, in order to maintain the channel closed in the resting state Citation[29]. When perturbed, SecE may relax this grip, enabling the channel to open. This would require the concomitant displacement of the ‘plug’ domain to a location at the periphery of the complex, in addition to a widening of the central pore to accommodate the substrate protein. These conformational changes are presumably directed by the association of the partner protein and substrate, and the subsequent forces brought to bear by protein synthesis or the hydrolytic cycle of SecA. The separation of the two halves of SecY, comparable to a ‘clam-shell’ also offers an escape route to the membrane. Therefore, the structure is compatible with the ability of the complex to facilitate the passage of proteins both through and into the membrane.

There are a few clues with respect to how the channel might open. The ‘plug’ has been shown to cross-link to the C-terminus of SecE, approximately 20 Å away from its position in the structure Citation[32], Citation[33], identifying a potential pocket which it can inhabit during protein translocation. The formation of this cross-link, as well as the deletion of the plug resulted in enhanced protein translocation Citation[32], indicating that the complex may be held in a partially open state. The removal of the plug domain was however not lethal Citation[34] indicating that this domain is not the only determinant for channel regulation. The Hsp70 homologue BiP involved in the eukaryotic post-translational translocation reaction is thought to contribute to the seal formed by the channel Citation[35]. However, prokaryotes lack this component, indicating that other mechanisms can be adopted to provide a barrier to small molecules. A careful comparison of the membrane bound dimer Citation[28] and the detergent solubilized monomer Citation[29], indicates that in the former the plug has moved about 6 Å towards the outside, and the lateral gate for membrane protein insertion has opened slightly () Citation[36]. In this context, although closed, the dimer of SecYEG appears to be primed for protein translocation.

Figure 2.  (A) Homology model of E. coli SecYEG based on the detergent solubilized SecYEβ monomer. (B) Model fitted according to the membrane bound dimer Citation[36]. Both are side views with the cytosolic face uppermost; the latter shows only the membrane sector, as the loops were not resolved sufficiently. TM2b (olive) and TM7 (green) form the signal sequence-binding site and have parted slightly in the dimer. The ‘plug’ (red) is obscured in (A) and is about 6.5 Å higher. For purposes of clarity schematic representations have been drawn on the right hand sides.

Figure 2.  (A) Homology model of E. coli SecYEG based on the detergent solubilized SecYEβ monomer. (B) Model fitted according to the membrane bound dimer Citation[36]. Both are side views with the cytosolic face uppermost; the latter shows only the membrane sector, as the loops were not resolved sufficiently. TM2b (olive) and TM7 (green) form the signal sequence-binding site and have parted slightly in the dimer. The ‘plug’ (red) is obscured in (A) and is about 6.5 Å higher. For purposes of clarity schematic representations have been drawn on the right hand sides.

One puzzling aspect of the structure is the fact that the monomer appears to provide all of the components for translocation: the signal sequence binding site, the protein-pore, exposed cytosolic loops for partner protein interaction and a lateral and traversal exit site. Several studies from independent laboratories have observed the bacterial channel in a dimeric form Citation[28], Citation[33], Citation[37–40]. The reasons for the presence of high oligomeric states are not entirely clear. The complex may exist as an oligomer with multiple independent active sites as many other soluble and membrane proteins do. Alternatively, the larger assemblies might be required to provide a large enough platform for the association of the more sizable partners. Specific inter-subunit interactions might also be required for activation or to support a dynamic process required for the reaction. Finally, the mechanism might rely on the partner proteins or substrate to introduce an asymmetric element important for the reaction, to generate a singly active protomer. A recent study has shown that protein translocation occurs through only one copy of the two that are present in the active complex (and not through a consolidated channel) [Osborne & Rapoport (2007) Cell, 129, 97–110].

The ‘back-to-back’ orientation observed of the membrane bound form has also been detected biochemically. In this particular form TM3 of SecE is located at the interface and the lateral gate for membrane protein insertion (between TM2 & TM7 of SecY; ) point in opposite directions and toward the lipid bilayer Citation[28], Citation[29], Citation[36]. Cysteine mutagenesis experiments identified residues in SecE that were close to the equivalent helix in the neighbouring monomer Citation[41], Citation[42]. These observations were enhanced by conditions that promote a productive association of the partner protein SecA Citation[41]. Moreover, the cross-linked dimer retained the capacity to tightly bind and activate SecA, but had lost the ability to couple ATP hydrolysis to the work of translocation Citation[41]. A careful inspection of the position of the respective side chains reveal that, although close, they point away from each another (). The association of SecA may have perturbed them and brought them closer together, and thus more amenable to cross-linking. The formation of a disulphide in this position would bring about a considerable distortion of the two helices. It is conceivable that this imposed distortion and restriction on the motility at the dimer interface may have had a minor affect on the association with SecA, but a disastrous one for the energy coupling process.

Figure 3.  Space filling representation of the dimeric membrane-bound E. coli atomic model Citation[36]. SecY and SecE are coloured green and red, respectively. The green arrows denote the position of the lateral gate. The E. coli cysteine cross-link (L106C) is shown in blue. The distance between the two cysteines is 18.5 Å, contrasting to the 2.05Å disulphide bond length. (A) View from the periplasmic face of the membrane. (B) View from the side of the membrane. Simplified views have been drawn underneath; the open blue circle (B) denotes that the cysteine is on the other side of TM3 of SecE.

Figure 3.  Space filling representation of the dimeric membrane-bound E. coli atomic model Citation[36]. SecY and SecE are coloured green and red, respectively. The green arrows denote the position of the lateral gate. The E. coli cysteine cross-link (L106C) is shown in blue. The distance between the two cysteines is 18.5 Å, contrasting to the 2.05Å disulphide bond length. (A) View from the periplasmic face of the membrane. (B) View from the side of the membrane. Simplified views have been drawn underneath; the open blue circle (B) denotes that the cysteine is on the other side of TM3 of SecE.

Another study by electron cryo-microscopy has revealed two SecYEG dimers with a ribosome-nascent chain complex (one bound close to the exit channel and the other to mRNA) Citation[39]. The map was of insufficient detail to resolve clear and individual TM domains, but could be used to fit the high-resolution structure into the density. On the basis of normal mode-based flexible fitting the ‘back-to-back’ SecYEG structure was rejected in favour of a new ‘front-to-front’ model. This arrangement places the same TM domain of SecE on the opposite side of the membrane complex, and the two lateral gates close to each other at the dimer interface. This proposal has been used as a foundation for a model of translocation incorporating both monomers of SecYEG in a single translocation cycle; in this context it has been suggested that the channels open at the dimer interface to form a consolidated environment for protein transport Citation[39], Citation[43], Citation[44]. However, disulfide cross-linking shows that the contacts the translocating pre-protein makes with SecY are within the strict confides of the channel, and not, as might be expected from such an arrangement, in the teeth of the ‘clam-shell’ Citation[31].

The structure and domain organization of SecA

Like other molecular motors SecA must undergo large ATP-dependent conformational changes and couple them to the domain movements required for function. The key step for energy transduction in SecA must be the conformational change regulated by ATP binding and product release in the nucleotide-binding fold (NBF). It is how these movements are relayed through the enzyme and to their partners, which provides the power stroke, and holds the key to understanding the reaction mechanism.

SecA is a soluble protein of 102 kDa and exists in solution in a monomer-dimer equilibrium, predominantly dimeric Citation[45], Citation[46]. Of the six published SecA crystal structures Citation[47–52], five of them are dimeric and one monomeric (). The protomers of each dimer all have a similar structure, the major differences between them being at the dimer interface. Interestingly, the structure from Thermus thermophilus is in a parallel conformation Citation[47], in contrast to all of the other structures which are packed in an anti-parallel manner Citation[48], Citation[50–52]. It is not clear which of these structures is the correct physiological state, as it is possible that the oligomeric arrangement observed in X-ray structures is a consequence of the extreme crystallization conditions or the lattice contacts. The parallel and two anti-parallel forms have however been shown to exist in solution by directed cysteine mutagenesis and cross-linking Citation[47], Citation[50] suggesting that in the absence of other translocation partners and pre-protein there may be different conformational states of the enzyme.

Figure 4.  Ribbon representation of the monomeric SecA structure Citation[49]. NBF1 is shown in dark green, NBF2 pale blue, HSD purple, HWD bright green, PPXD in dark blue and Mg2 + -ADP in orange/red. The arrow denotes the position of the putative polypeptide-binding site.

Figure 4.  Ribbon representation of the monomeric SecA structure Citation[49]. NBF1 is shown in dark green, NBF2 pale blue, HSD purple, HWD bright green, PPXD in dark blue and Mg2 + -ADP in orange/red. The arrow denotes the position of the putative polypeptide-binding site.

Within each protomer there are five major domains, from N- to C-terminus these are: nucleotide binding fold 1 (NBF1), pre-protein cross-linking domain (PPXD) Citation[53], Citation[54], nucleotide binding fold 2 (NBF2), helical scaffold domain (HSD) and helical wing domain (HWD) Citation[51]. In the monomeric crystal, movements of the HWD, HSD and PPXD domains results in opening of a groove that may be the polypeptide binding site Citation[49], Citation[55] (). In all of the apo- and nucleotide-bound structures, there are essentially no major changes in the structure of the NBFs Citation[48], Citation[51], Citation[56]. This means that there are missing active states of the complex. The constraints of the crystallization process or the transient nature of these states may mean that they are difficult to characterise structurally.

Interaction sites and stoichiometry of SecA and SecYEG

SecA has been shown to associate with membranes containing SecYEG to a high affinity Citation[8]. The structure of the SecYEG complex indicates that there are loops exposed to the cytosol available for partner protein interaction Citation[28], Citation[29], Citation[36], but the precise locations are ill defined. Both the N-terminal domain (comprising the NBF) and C-terminal domains of SecA may be important for the interaction with SecYEG Citation[57–59]. Second site suppressors, cross-linking and peptide scanning experiments map the interaction sites in SecY to TM4, the fourth and fifth cytoplasmic loops (C4 and C5) and the C-terminal tail Citation[60–63]. There is evidence to support the fact that the dimer of SecA is an active form. The inactivation of a single SecA protomer has a dominant effect on the activity of dimer and resonance energy transfer experiments designed to monitor the appearance of monomers proved negative Citation[64]. However, recent reports have challenged this view. The fact that one of the structures of SecA is a monomer led to suggestion that the dimer might dissociate when activated Citation[49], Citation[50]. The observed loss of specific inter-molecular cross-links and fluorescence resonance energy transfer between dimeric SecA on exposure to translocation specific ligands is an indication of monomerization Citation[49], Citation[65], Citation[66]. These observations may also be the result of a rearrangement to form one of the other dimeric structures, or a hitherto uncharacterized and active dimeric form of the protein. During translocation, a SecA monomer remained associated with a SecYEG protomer and the pre-protein Citation[67]. Two further studies identified that both monomers and dimers of SecA were able to associate with SecYEG dimers in solution, and that the specific oligomeric association was dependent on nucleotide Citation[37], Citation[67]. Perhaps both monomers and dimers of SecA play important roles in the translocation reaction, and association and dissociation are required to complete the reaction cycle.

The hydrolytic cycle of SecA

The ATPase activity of SecA is sensitive to ligands, including: mature proteins Citation[68], signal peptides Citation[68–71], SecB Citation[72], Citation[73], ADP Citation[46], magnesium Citation[46], lipids Citation[8], Citation[46], Citation[65], Citation[66], Citation[68], Citation[74–76] and SecYEG Citation[8], Citation[68]. Two specific regulatory domains for ATPase activity have been proposed. The first Intramolecular Regulator of ATPase activity (IRA1) resides at the far C-terminus and controls ATP hydrolysis in both soluble and SecY-associated SecA Citation[77]. The second, IRA2 (NBF2) controls the nucleotide binding and release occurring at NBF1 Citation[78].

The ATP bound state of SecA has been shown to be in an extended conformation Citation[79], whereas the ADP bound form is more compact Citation[51], Citation[77–81], the release of which is thought to be the rate limiting step of the steady-state reaction cycle Citation[80]. The non-hydrolysable ATP analogue AMP-PNP stabilises the association of SecA and SecYEG and promotes the ‘inserted’ state of a substrate protein Citation[37], Citation[82]. Experiments employing surface plasmon resonance and cross-linking show that ATP induces the release of SecA from SecYEG Citation[67], Citation[83].

The divalent cation magnesium exerts a strong inhibitory effect upon SecA, affecting both the dimeric conformation (but not oligomeric state) and the affinity for ATP; it does this by acting on an allosteric binding site distinct from the catalytic one Citation[46]. This inhibition could be overcome by the acidic phospholipid cardiolipin, which had a stimulatory role only in the presence of magnesium. Lipid binding sites have been found on SecA Citation[84], Citation[85], and acidic phospholipids are required for translocation Citation[86]. This suggests that the ‘lipid-ATPase’ activity reported may in fact be related to a counteraction of magnesium inhibition by cardiolipin-SecA interactions Citation[46]. This inhibition presumably prevents the futile hydrolysis of ATP during moments of inactivity, and is released as a result of intimate exposure to the membrane.

The catalytic cycle has been proposed to proceed by a multi-step mechanism for translocation via SecA insertion and de-insertion events Citation[82]. The evidence is based upon the protease protection of a large domain of SecA while bound by SecY, which is dependent on ATP and pre-protein Citation[82], Citation[87]. This protection is the result of a necessary and uncharacterized large conformational change of the peripherally associated protein and not necessarily the result of a deep penetration of the enzyme. The thermodynamics of this reaction have also not been well characterized; one study proposes that one ATP drives the passage of 40 amino acids, and another determines that each amino acid requires 5 molecules of ATP Citation[88], Citation[89].

Concluding comments

This is an extremely complicated field of membrane biology and our comprehension of it is thanks to a generation of research on the related eukaryotic, bacterial and archaeal systems. In spite of the recent progress there are several disputed and unanswered aspects of the reaction mechanism. These outstanding problems will need to be addressed by detailed structural determination of an active state of the translocation complex with pre-protein, and by further biophysical studies aimed to comprehend the dynamics of the reaction. Understanding details both on the nature and timing of the conformational changes that derive from the release of chemical energy, and of specific protein-protein interactions holds the key to understanding how they are coupled to protein transport.

References

  • Walter P, Blobel G. Translocation of proteins across the endoplasmic reticulum. III. Signal recognition protein (SRP) causes signal sequence and site specific arrest of chain elongation that is released by microsomal membranes. J Cell Biol 1981; 91: 557–561
  • Walter P, Ibrahimi I, Blobel G. Translocation of proteins across the endoplasmic reticulum. I. Signal recognition protein (SRP) binds to in vitro assembled polysomes synthesizing secretory protein. J Cell Biol 1981; 91: 545–550
  • Gilmore R, Blobel G. Transient involvement of signal recognition particle and its receptor in the microsomal membrane prior to protein translocation. Cell 1983; 35: 677–685
  • Beckmann R, Spahn C, Eswar N, Helmers J, Penczek P, Sali A, Frank J, Blobel G. Architecture of the protein-conducting channel associated with the translating 80S ribosome. Cell 2001; 107: 361–372
  • Görlich D, Prehn S, Hartmann E, Kalies K, Rapoport T. A mammalian homolog of SEC61p and SECYp is associated with ribosomes and nascent polypeptides during translocation. Cell 1992; 71: 489–503
  • Misselwitz B, Staeck O, Matlack K, Rapoport T. Interaction of BiP with the J-domain of the Sec63p component of the endoplasmic reticulum protein translocation complex. J Biol Chem 1999; 274: 20110–20115
  • Matlack K, Misselwitz B, Plath K, Rapoport T. BiP acts as a molecular ratchet during posttranslational transport of prepro-alpha factor across the ER membrane. Cell 1999; 97: 553–564
  • Hartl F, Lecker S, Schiebel E, Hendrick J, Wickner W. The binding cascade of SecB to SecA to SecY/E mediates preprotein targeting to the E. coli plasma membrane. Cell 1990; 63: 269–279
  • Brundage L, Hendrick JP, Schiebel E, Driessen AJM, Wickner W. The purified E. Coli integral membrane protein SecY/E is sufficient for reconstitution of SecA-dependent precursor protein translocation. Cell 1990; 62: 649–657
  • Weiss J, Ray P, Bassford P. Purified SecB protein of Escherichia coli retards folding and promotes membrane translocation of the maltose-binding protein in vivo. Proc Natl Acad Sci USA 1988; 85: 8978–8982
  • Kusters R, de Vrije T, Breukink E, de Kruijff B. SecB protein stabilizes a translocation-competent state of purified PrePhoe protein. J Biol Chem 1989; 264: 20827–20830
  • de Keyzer J, van der Does C, Driessen AJ. Kinetic analysis of the translocation of fluorescent precursor proteins into Escherichia coli membrane vesicles. J Biol Chem 2002; 277: 46059–46065
  • Kumamoto CA, Beckwith J. Mutations in a new gene, secB, cause defective protein localization in Escherichia coli. J Bacteriol 1983; 154: 253–260
  • Oliver D, Beckwith J. E. coli mutant pleiotropically defective in the export of secreted proteins. Cell 1981; 25: 765–772
  • Oliver D, Beckwith J. Identification of a new gene (secA) and gene product involved in the secretion of envelope proteins in Escherichia coli. J Bacteriol 1982; 150: 686–691
  • Emr SD, Hanley-Way S, Silhavy T. Suppressor mutations that restore export of a protein with a defective signal sequence. Cell 1981; 23: 79–88
  • Brickman E, Oliver D, Garwin J, Kumamoto C, Beckwith J. The use of extragenic suppressors to define genes involved in protein export in Escherichia coli. Mol Gen Genet 1984; 196: 24–27
  • Oliver DB, Liss LR. prlA-mediated suppression of signal sequence mutations is modulated by the secA gene product of Escherichia coli K-12. J Bacteriol 1985; 161: 817–819
  • Chen L, Tai PC. ATP is essential for protein translocation into Escherichia coli membrane vesicles. Proc Natl Acad Sci USA 1985; 82: 4384–4388
  • Lill R, Cunningham K, Brundage L, Ito K, Oliver D, Wickner W. SecA protein hydrolyzes ATP and is an essential component of the protein translocation ATPase of Escherichia coli. EMBO J 1989; 8: 961–966
  • Bieker KL, Silhavy TJ. PrlA (SecY) and PrlG (SecE) Interact directly and function sequentially during protein translocation in E. Coli. Cell 1990; 61: 833–842
  • Hanada M, Nishiyama K, Mizushima S, Tokuda H. Reconstitution of an efficient protein translocation machinery comprising SecA and the three membrane proteins, SecY, SecE, and SecG (p12). J Biol Chem 1994; 269: 23625–23631
  • Hanein D, Matlack K, Jungnickel B, Plath K, Kalies K, Miller K, Rapoport T, Akey C. Oligomeric rings of the Sec61p complex induced by ligands required for protein translocation. Cell 1996; 87: 721–732
  • Beckmann R, Bubeck D, Grassucci R, Penczek P, Verschoor A, Blobel G, Frank J. Alignment of conduits for the nascent polypeptide chain in the ribosome-Sec61 complex. Science 1997; 278: 2123–2126
  • Menetret J, Neuhof A, Morgan D, Plath K, Radermacher M, Rapoport T, Akey C. The structure of ribosome-channel complexes engaged in protein translocation. Mol Cell 2000; 6: 1219–1232
  • Morgan DG, Menetret J-F, Neuhof A, Rapoport TA, Akey CW. Structure of the mammalian Ribosome-Channel complex at 17 Å resolution. J Mol Biol 2002; 324: 871–886
  • Collinson I, Breyton C, Duong F, Tziatzios C, Schubert D, Or E, Rapoport T, Kuhlbrandt W. Projection structure and oligomeric properties of a bacterial core protein translocase. EMBO J 2001; 20: 2462–2471
  • Breyton C, Haase W, Rapoport TA, Kuhlbrandt W, Collinson I. Three-dimensional structure of the bacterial protein-translocation complex SecYEG. Nature 2002; 418: 662–665
  • van den Berg B, Clemons WM, Jr, Collinson I, Modis Y, Hartmann E, Harrison SC, Rapoport TA. X-ray structure of a protein-conducting channel. Nature 2004; 427: 36–44
  • Plath K, Mothes W, Wilkinson B, Stirling C, Rapoport T. Signal sequence recognition in posttranslational protein transport across the yeast ER membrane. Cell 1998; 94: 795–807
  • Cannon KS, Or E, Clemons WM, Jr, Shibata Y, Rapoport TA. Disulfide bridge formation between SecY and a translocating polypeptide localizes the translocation pore to the center of SecY”. J Cell Biol” 2005; 169: 219–225
  • Harris CR, Silhavy TJ. Mapping an interface of SecY (PrlA) and SecE (PrlG) by using synthetic phenotypes and in vivo cross-linking. J Bacteriol 1999; 181: 3438–3444
  • Tam PC, Maillard AP, Chan KK, Duong F. Investigating the SecY plug movement at the SecYEG translocation channel. EMBO J 2005; 24: 3380–3388
  • Maillard AP, Lalani S, Silva F, Belin D, Duong F. Deregulation of the SecYEG translocation channel upon removal of the plug domain. J Biol Chem 2007; 282: 1281–1287
  • Alder NN, Shen Y, Brodsky JL, Hendershot LM, Johnson AE. The molecular mechanisms underlying BiP-mediated gating of the Sec61 translocon of the endoplasmic reticulum. J Cell Biol 2005; 168: 389–399
  • Bostina M, Mohsin B, Kuhlbrandt W, Collinson I. Atomic model of the E. coli membrane-bound protein translocation complex SecYEG. J Mol Biol 2005; 352: 1035–1043
  • Tziatzios C, Schubert D, Lotz M, Gundogan D, Betz H, Schagger H, Haase W, Duong F, Collinson I. The bacterial protein-translocation complex: SecYEG dimers associate with one or two SecA molecules. J Mol Biol 2004; 340: 513–524
  • Bessonneau P, Besson V, Collinson I, Duong F. The SecYEG preprotein translocation channel is a conformationally dynamic and dimeric structure. EMBO J 2002; 21: 995–1003
  • Mitra K, Schaffitzel C, Shaikh T, Tama F, Jenni S, Brooks CL 3rd, Ban N, Frank J. Structure of the E. coli protein-conducting channel bound to a translating ribosome. Nature 2005; 438: 318–324
  • Scheuring J, Braun N, Nothdurft L, Stumpf M, Veenendaal AK, Kol S, van der Does C, Driessen AJ, Weinkauf S. The oligomeric distribution of SecYEG is altered by SecA and translocation ligands. J Mol Biol 2005; 354: 258–271
  • Kaufmann A, Manting EH, Veenendaal AK, Driessen AJ, van der Does C. Cysteine-directed cross-linking demonstrates that helix 3 of SecE is close to helix 2 of SecY and helix 3 of a neighboring SecE. Biochemistry 1999; 38: 9115–9125
  • Veenendaal A, van der Does C, Driessen A. Mapping the sites of interaction between SecY and SecE by cysteine scanning mutagenesis. J Biol Chem 2001; 276: 32559–32566
  • Mitra K, Frank J. A model for co-translational translocation: Ribosome-regulated nascent polypeptide translocation at the protein-conducting channel. FEBS Lett 2006; 580: 3353–3360
  • Mitra K, Frank J, Driessen A. Co- and post-translational translocation through the protein-conducting channel: analogous mechanisms at work?. Nat Struct Mol Biol 2006; 13: 957–964
  • Woodbury RL, Hardy SJS, Randall LL. Complex behavior in solution of homodimeric SecA. Protein Sci 2002; 11: 875–882
  • Gold, VAM, Robson, A, Clarke, AR, Collinson, I. 2007. Allosteric regulation of SecA: magnesium-mediated control of conformation and activity. J Biol Chem ( in press).
  • Vassylyev DG, Mori H, Vassylyeva MN, Tsukazaki T, Kimura Y, Tahirov TH, Ito K. Crystal structure of the translocation ATPase SecA from Thermus thermophilus reveals a parallel, head-to-head dimer. J Mol Biol 2006; 364: 248–258
  • Papanikolau Y, Papadovasilaki M, Ravelli RB, McCarthy AA, Cusack S, Economou A, Petratos K. Structure of dimeric SecA, the Escherichia coli preprotein translocase motor. J Mol Biol 2006; 366: 1545–1557
  • Osborne AR, Clemons WM, Jr, Rapoport TA. A large conformational change of the translocation ATPase SecA. Proc Natl Acad Sci USA 2004; 101: 10937–10942
  • Zimmer J, Li W, Rapoport TA. A novel dimer interface and conformational changes revealed by an X-ray structure of B. subtilis SecA. J Mol Biol 2006; 364: 259–265
  • Hunt JF, Weinkauf S, Henry L, Fak JJ, McNicholas P, Oliver DB, Deisenhofer J. Nucleotide control of interdomain interactions in the conformational reaction cycle of SecA. Science 2002; 297: 2018–2026
  • Sharma V, Arockiasamy A, Ronning DR, Savva CG, Holzenburg A, Braunstein M, Jacobs WR, Jr, Sacchettini JC. Crystal structure of Mycobacterium tuberculosis SecA, a preprotein translocating ATPase. Proc Natl Acad Sci USA 2003; 100: 2243–2248
  • Kimura E, Akita M, Matsuyama S, Mizushima S. Determination of a region in SecA that interacts with presecretory proteins in Escherichia coli. J Biol Chem 1991; 266: 6600–6606
  • Kourtz L, Oliver D. Tyr-326 plays a critical role in controlling SecA-preprotein interaction. Mol Microbiol 2000; 37: 1342–1356
  • Musial-Siwek M, Rusch SL, Kendall DA. Selective photoaffinity labeling identifies the signal peptide binding domain on SecA. J Mol Biol 2007; 365: 637–648
  • Osborne AR, Rapoport TA, van den Berg B. Protein translocation by the Sec61/SecY channel. Annu Rev Cell Dev Biol 2005; 21: 529–550
  • Vrontou E, Karamanou S, Baud C, Sianidis G, Economou A. Global co-ordination of protein translocation by the SecA IRA1 switch. J Biol Chem 2004; 279: 22490–22497
  • Matsumoto G, Nakatogawa H, Mori H, Ito K. Genetic dissection of SecA: suppressor mutations against the secY205 translocase defect. Genes Cells 2000; 5: 991–999
  • Snyders S, Ramamurthy V, Oliver D. Identification of a region of interaction between Escherichia coli SecA and SecY proteins. J Biol Chem 1997; 272: 11302–11306
  • Taura T, Akiyama Y, Ito K. Genetic analysis of SecY: additional export-deficient mutations and factors affecting their phenotypes. Mol Gen Genet 1994; 243: 261–269
  • Mori H, Ito K. Biochemical characterization of a mutationally altered protein translocase: proton motive force stimulation of the initiation phase of translocation. J Bacteriol 2003; 185: 405–412
  • van der Sluis EO, Nouwen N, Koch J, de Keyzer J, van der Does C, Tampe R, Driessen AJ. Identification of two interaction sites in SecY that are important for the functional interaction with SecA. J Mol Biol 2006; 361: 839–849
  • Mori H, Ito K. Different modes of SecY-SecA interactions revealed by site-directed in vivo photo-cross-linking. Proc Natl Acad Sci USA 2006; 103: 16159–16164
  • Driessen A. SecA, the peripheral subunit of the Escherichia coli precursor protein translocase, is functional as a dimer. Biochemistry 1993; 32: 13190–13197
  • Or E, Navon A, Rapoport TA. Dissociation of the dimeric SecA ATPase during protein translocation across the bacterial membrane. EMBO J 2002; 21: 4470–4479
  • Or E, Boyd D, Gon S, Beckwith J, Rapoport T. The bacterial ATPase SecA functions as a monomer in protein translocation. J Biol Chem 2005; 280: 9097–9105
  • Duong F. Binding, activation and dissociation of the dimeric SecA ATPase at the dimeric SecYEG translocase. EMBO J 2003; 22: 4375–4384
  • Lill R, Dowhan W, Wickner W. The ATPase activity of SecA is regulated by acidic phospholipids, SecY, and the leader and mature domains of precursor proteins. Cell 1990; 60: 271–280
  • Wang L, Miller A, Kendall DA. Signal peptide determinants of SecA binding and stimulation of ATPase activity. J Biol Chem 2000; 275: 10154–10159
  • Triplett TL, Sgrignoli AR, Gao FB, Yang YB, Tai PC, Gierasch LM. Functional signal peptides bind a soluble N-terminal fragment of SecA and inhibit its ATPase activity. J Biol Chem 2001; 276: 19648–19655
  • Miller A, Wang L, Kendall DA. Synthetic signal peptides specifically recognize SecA and stimulate ATPase activity in the absence of preprotein. J Biol Chem 1998; 273: 11409–11412
  • Miller A, Wang L, Kendall DA. SecB modulates the nucleotide-bound state of SecA and stimulates ATPase activity. Biochemistry 2002; 41: 5325–5332
  • Kim J, Miller A, Wang L, Muller JP, Kendall DA. Evidence that SecB enhances the activity of SecA. Biochemistry 2001; 40: 3674–3680
  • Benach J, Chou Y-T, Fak JJ, Itkin A, Nicolae DD, Smith PC, Wittrock G, Floyd DL, Golsaz CM, Gierasch LM, Hunt JF. Phospholipid-induced monomerization and signal-peptide-induced oligomerization of SecA. J Biol Chem 2003; 278: 3628–3638
  • Bu Z, Wang L, Kendall DA. Nucleotide binding induces changes in the oligomeric state and conformation of SecA in a lipid environment: A small-angle neutron-scattering study. J Mol Biol 2003; 332: 23–30
  • Ahn T, Kim H. Effects of nonlamellar-prone lipids on the ATPase activity of SecA bound to model membranes. J Biol Chem 1998; 273: 21692–21698
  • Karamanou S, Vrontou E, Sianidis G, Baud C, Roos T, Kuhn A, Politou AS, Economou A. A molecular switch in SecA protein couples ATP hydrolysis to protein translocation. Mol Microbiol 1999; 34: 1133–1145
  • Sianidis G, Karamanou S, Vrontou E, Boulias K, Repanas K, Kyrpides N, Politou AS, Economou A. Cross-talk between catalytic and regulatory elements in a DEAD motor domain is essential for SecA function. EMBO J 2001; 20: 961–970
  • den Blaauwen T, Fekkes P, de Wit JG, Kuiper W, Driessen AJ. Domain interactions of the peripheral preprotein Translocase subunit SecA. Biochemistry 1996; 35: 11994–12004
  • Fak JJ, Itkin A, Ciobanu DD, Lin EC, Song XJ, Chou YT, Gierasch LM, Hunt JF. Nucleotide exchange from the high-affinity ATP-binding site in SecA is the rate-limiting step in the ATPase cycle of the soluble enzyme and occurs through a specialized conformational state. Biochemistry 2004; 43: 7307–7327
  • Kim J, Ahn T, Ko J, Park C, Kim H. Effect of divalent cations on the ATPase activity of Escherichia coli SecA. FEBS Lett 2001; 493: 12–16
  • Economou A, Wickner W. SecA promotes preprotein translocation by undergoing ATP-driven cycles of membrane insertion and deinsertion. Cell 1994; 78: 835–843
  • de Keyzer J, van der Does C, Kloosterman TG, Driessen AJ. M. Direct demonstration of ATP-dependent release of SecA from a translocating preprotein by surface plasmon resonance. J Biol Chem 2003; 278: 29581–29586
  • Breukink E, Nouwen N, van Raalte A, Mizushima S, Tommassen J, de Kruijff B. The C terminus of SecA is involved in both lipid binding and SecB binding. J Biol Chem 1995; 270: 7902–7907
  • Breukink E, Keller RCA, de Kruijff B. Nucleotide and negatively charged lipid-dependent vesicle aggregation caused by SecA. Evidence that SecA contains two lipid-binding sites. FEBS Lett 1993; 331: 19–24
  • Hendrick JP, Wickner W. SecA protein needs both acidic phospholipids and SecY/E protein for functional high-affinity binding to the Escherichia coli plasma membrane. J Biol Chem 1991; 266: 24596–24600
  • Jilaveanu LB, Oliver DB. In vivo membrane topology of Escherichia coli SecA ATPase reveals extensive periplasmic exposure of multiple functionally important domains clustering on one face of SecA. J Biol Chem 2007; 282: 4661–4668
  • Tomkiewicz D, Nouwen N, van Leeuwen R, Tans S, Driessen AJ. SecA supports a constant rate of preprotein translocation. J Biol Chem 2006; 281: 15709–15713
  • van der Wolk J, de Wit J, Driessen A. The catalytic cycle of the Escherichia coli SecA ATPase comprises two distinct preprotein translocation events. EMBO J 1997; 16: 7297–7304

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.