doi:10.1016/S0022-2836(03)00840-4 J. Mol. Biol. (2003) 332, 23–30

COMMUNICATION Nucleotide Binding Induces Changes in the Oligomeric State and Conformation of Sec A in a Lipid Environment: A Small-angle Neutron-scattering Study

Zimei Bu, Ligong Wang and Debra A. Kendall*

Department of Molecular In Escherichia coli, SecA is a large, multifunctional protein that is a vital and Cell Biology component of the general protein secretion pathway. In its membrane- University of Connecticut bound form it functions as the motor component of the protein trans- 91 N. Eagleville Road locase, perhaps through successive rounds of membrane insertion and Storrs, CT 06269-3125 ATP hydrolysis. To understand both the energy conversion process USA and assembly, we have used contrast-matched, small-angle neutron-scattering (SANS) experiments to examine SecA in small uni- lamellar vesicles of E. coli phospholipids. In the absence of nucleotide, we observe a dimeric form of SecA with a radius of gyration comparable to that previously observed for SecA in solution. In contrast, the presence of either ADP or a non-hydrolyzable ATP analog induces conversion to a monomeric form. The larger radius of gyration for the ATP-bound relative to the ADP-bound form suggests the former has a more expanded global conformation. This is the first direct structural determination of SecA in a lipid bilayer. The SANS data indicate that nucleotide turnover can func- tion as a switch of conformation of SecA in the membrane in a manner consistent with its proposed role in successive cycles of deep membrane penetration and release with concommitant preprotein insertion. q 2003 Elsevier Ltd. All rights reserved. Keywords: SecA; oligomerization; nucleotide; protein transport; small- *Corresponding author angle neutron scattering

Many proteins that are synthesized in the cyto- associated forms,6 and interfaces with SecYEG and plasm of cells use the Sec translocase for routing the preprotein to propel preprotein translocation to non-cytoplasmic locations. In Escherichia coli, at the expense of ATP. Biochemical studies suggest SecYEG and SecA are critical components of this that upon binding ATP7 and lipid,8 SecA under- translocase that function in concert with the pre- goes a substantial conformational change9 and is protein to ensure its passage through the mem- deeply inserted into the membrane bilayer. brane. SecYEG can be isolated and purified as a This membrane association results in stimulation stable heterotrimeric complex,1–5 and constitutes of SecA ATPase activity and preprotein trans- the central component of the membrane-embedded location.10 – 12 translocase. SecA is a large, multifunctional protein A fundamental issue to be resolved is how SecA that is found in both cytoplasmic and membrane- translates the chemical energy of ATP hydrolysis into the mechanical energy required for preprotein movement. To understand the mechanistic basis Present address: Z. Bu, Fox Chase Cancer Center, for this activity, the structural changes of mem- Philadelphia, PA 19111, USA. 0 brane-associated SecA during nucleotide turnover Abbreviations used: AMP-PNP, 5 -adenylylimidodi- must be elucidated. This is a complex problem to phosphate; BSA, bovine serum albumin; SUV, small study, however, because the requirement for the unilamellar vesicles; SANS, small-angle neutron- scattering; SAXS, small angle X-ray scattering; SUV, membrane-bound form renders many classical small unilamellar vesicles. methods of structural analysis of proteins in E-mail address of the corresponding author: aqueous solution unfeasible. The SecA-nucleotide [email protected] cycling event is a dynamic process, making studies

0022-2836/$ - see front matter q 2003 Elsevier Ltd. All rights reserved. 24 Nucleotide-induced SecA Conformational Changes

Figure 1. (a) Contrast variation experiments to determine the match point of E. coli SUV. To make SUV, 0.4 ml of 25 mg/ml of chloro- form-dissolved E. coli polar lipid extract (Avanti Polar Lipids, Inc., AL) was transferred to a 12 ml Pyrex tube. A thin film of lipid was formed under a stream of filtered nitrogen and residual chloroform was removed by high vacuum overnight. The lipid film was then re-hydrated in 1 mM DTT, 50 mM potassium phosphate (pH 7.5), 50 mM KCl, 1 mM EDTA. SUV were formed by sonicating until the solution was optically clear. The contrast variation experiments required different volume fractions 2 of H2O to determine the lipid contrast match point at which the scattering intensity from the SUV is zero. For contrast variation experi- ments, the lipid was resuspended 2 separately in H2O and H2O buffer to make the two SUV solutions. 2 The H2O and H2O SUV solutions were mixed in different proportions to produce six SUV solutions of 2 different H2O volume fractions. The lipid concentration used for contrast variation experiments was 10 mg/ml. The square-root of scattering intensity I(0)0.5, from the 2 SUV as a function of H2O volume 2 fraction is shown. The contrast match point can be determined from this plot to be X2H2O ¼ 0.143. At a H2O volume fraction of 0.143, the scattering intensity from the lipid vesicles is zero, indicating the scattering density contrast between the SUV and the buffer solution is zero. Further experiments on SecA/SUV solutions were therefore con- 2 ducted at this H2O concentration to eliminate the scattering from the SUV. (b) Association of SecA with SUV analyzed by flotation gradient centrifugation. SecA (1.5 mg/ml) was mixed with E. coli SUV (12 mg/ml) in the aforementioned phosphate buffer with density of 1.29 g/ml adjusted with metrizamide. Magnesium acetate (60 mM) was added when ADP (60 mM) or AMP-PNP (2 mM) was included in buffer. Samples (200 ml) were layered below 400 mlof corresponding buffers with density of 1.27 g/ml. Samples were centrifuged (for 33 hours, at 4 8C, at 35,000 rpm) in a Beckman SW50.1 rotor with adapters for 5 £ 41 mm Beckman Ultraclear centrifuge tubes, six 100 ml fractions were harvested and the top three and bottom two fractions analyzed by SDS-PAGE and staining with Coomassie brilliant blue. in model systems useful for evaluating the species tivity, differential scanning calorimetry and at different stages of nucleotide turnover, yet any dynamic light-scattering suggested that SecA in such system needs to be physiologically relevant solution adopts different conformations, depend- in terms of the content and concentration of ing on the nature of nucleotide bound.9,20 Measure- components. ment of the molar ratio of bound ADP to SecA in Several studies of SecA in aqueous solution indi- solution indicates that only one high-affinity cate that a dimeric form predominates in the nucleotide- exists per SecA dimer.21 presence and in the absence of nucleotide,13 – 16 This and a study involving heterodimers that though conditions of lower temperature and include ATPase inactive subunits15 suggest that higher ionic strength favor a shift toward SecA maintains its dimeric form during trans- monomer.17 The recently reported X-ray structures location, yet recent analysis suggests that the of Bacillus subtilis SecA at 2.7 A˚ resolution18 and association of SecA with negatively charged Micrococcus tuberculosis SecA at 2.8 A˚ resolution19 phospholipids22 or detergents23 shifts the equi- also suggests that in aqueous media it forms a librium toward the monomeric form. The compli- dimer that does not undergo a substantial confor- cated nature of the analyses is illustrated by mational change upon nucleotide binding, con- evidence that the oligomeric state of the SecYE sistent with small angle X-ray scattering (SAXS) components of the translocase involves a analysis of SecA, also in an aqueous environment.14 monomer,24 dimer,25 and higher-order structure.26 However, earlier studies involving protease sensi- Here, we report for the first time the direct Nucleotide-induced SecA Conformational Changes 25

Figure 2. (a) Guinier plots of SecA, SecA–ATP-Mg2þ and SecA–ADP-Mg2þ in E. coli SUV. E. coli SecA was isolated from strain BL21.14 pCS1 as described7 with greater than 98% purity as assessed by SDS-PAGE, and exhibited character- istic levels of ATPase activity in the presence of lipid.27 The protein was concentrated to about 20 mg/ml using a 2 ml Centricon with a cut-off of 10 kDa. The protein was then dialyzed against 200 ml of sterilized 50 mM potassium phos- 2 phate (pH 7.5), 50 mM KCl, 1 mM EDTA containing 14.3% (v/v) H2O. This concentration of EDTA was chosen carefully to preclude the possible magnesium-induced liposome aggregation, while ensuring the occupancy of the high-affinity nucleotide-binding site and, under comparable conditions, SecA exhibits ATPase activity. SecA was added to the SUV solution resulting in a concentration of SecA of 2.9 mg/ml and 23 mg/ml of lipid. Under similar conditions, SecA associates stably with vesicles (Figure 1(b)). For the ATP-binding experiment, 2 mM non-hydrolyzable ATP analog (AMP-PNP, 50-adenylyl imidodiphosphate) and 60 mMMg2þ were added to the SecA/SUV solution. For the ADP-bind- ing experiment, 60 mM ADP and 60 mMMg2þ were added to the SecA/SUV solution. SANS measurements were con- ducted using the NG-7 30 meter SANS instrument at the National Institute of Standards and Technology (Gaithersburg, MD).42 The incident neutron wavelength was l ¼ 6A˚ with a wavelength resolution of Dl/l ¼ 11%. The sample-to-detector distances were 7 m and 1.5 m to give a Q range from 0.006 A˚ 21 to 0.33 A˚ 21. Quartz banjo cells of 1 mm path-length were used. The scattered intensity was placed on an absolute level using a calibrated secondary stan- dard and circularly averaged to yield the scattered intensity, I(Q), as a function of Q. The scattering from the buffer background was measured, corrected by the volume fraction displaced by the protein, and subtracted from the scatter- ing by the protein solution. Data acquisition times were eight hours and 20 minutes at sample-to-detector distances of 7 m and 1.5 m, respectively, for both protein solutions and buffer backgrounds. SANS measurements were performed at room temperature. For a multi-component system, such as a protein/lipid complex, the scattering intensity is:43 2 0 0 2 0 IðQÞ¼ðrp 2 rsÞ IpðQÞþðrp 2 rsÞðrL 2 rsÞIpLðQÞþðrL 2 rsÞ ILðQÞð1Þ 0 0 where rL is the averaged scattering density of the lipid vesicles, ILðQÞ is the scattering from the lipid vesicles, IpLðQÞ is the cross-term reflecting the interference between the scattering amplitudes of the protein and the lipid vesicles. The experiments were conducted at the contrast match point of the lipid so that the scattering intensity from the lipid is zero, and equation (1) is reduced to: 2 0 IðQÞ¼ðrp 2 rsÞ I ðQÞð2Þ 2 where the scattering intensity, IðQÞ; is proportional to the square of the contrast ðrp 2 rsÞ ; where rp is the scattering density of the protein and rs is the scattering density of the buffer. The radius of gyration, Rg, is related to the size and conformation of the protein and can be obtained from the Guinier approximation: 1 ln IðQÞ¼ln Ið0Þ 2 R2Q2 ð3Þ 3 g by linear least-squares fitting in the QRg # 1 region, where Ið0Þ is the forward scattering intensity that is related to the molecular mass of the macromolecule:

2 Mw 2 Ið0Þ¼ðrp 2 rsÞ c vp ð4Þ NA where c is the protein concentration, Mw is the protein molecular mass, NA is Avogadro’s number and vpis the partial specific volume of the protein. Bovine serum albumin (BSA) was used as the molecular mass standard and was dissolved at 11.4 mg/ml in the SUV solution under the conditions used for SecA. The scattering from BSA in the lipid vesicles was used as a control for molecular mass calibration. The scattering intensities of the different forms of SecA and that of BSA were normalized by the protein concentrations. (b) PðrÞ of SecA, SecA–ATP-Mg2þ, SecA–ADP-Mg2þ. Inverse Fourier transformation of IðQÞ gives the distance distribution function, PðrÞ; which is the probable frequency distribution of the vector length between scattering points in the macromolecules. The Dmax at which PðrÞ becomes zero reflects the maximum dimension of the macromolecule averaged over all orientations. The shape and the Dmax of the PðrÞ can be used to deduce the macromolecular shape or the distance between two domains within the protein. Both Rg and Ið0Þ can be determined from PðrÞ as the second and zeroth moments of PðrÞ: The Q range used to compute PðrÞ of the SecA dimer in lipid vesicles is 0:0164 , Q , 0:1A˚ 21 and 0:0164 , Q , 0:12 A˚ 21 for the nucleotide-bound form of SecA. In this case, the sampling theorem is still satisfied. 26 Nucleotide-induced SecA Conformational Changes structural measurement of the size and shape of Table 1. A summary of the size and oligomeric states of SecA in vesicles composed of E. coli phospholipids SecA measured by SANS to model the biological membrane. To distinguish Mw Rg ˚ SecA features from those of the vesicle, our I(0)/cprotein (kDa) (A) strategy employed contrast-matched, small-angle, ^ ^ neutron-scattering (SANS) to evaluate SecA in the Apo SecA 0.033 0.002 204 45.1 0.9 SecA þ Mg2þ þ ATP analog 0.017 ^ 0.001 105 38.3 ^ 0.8 absence and in the presence of nucleotide. We SecA þ Mg2þ þ ADP 0.019 ^ 0.001 117 31.1 ^ 0.5 find that SecA without nucleotide exists in a pre- BSA 0.011 ^ 0.001 68 31.6 ^ 0.2 dominantly dimeric form in this lipid environ- The error in protein concentration determination is about ment that is converted almost exclusively to 10%. The error in molecular mass (calibrated against BSA) is monomer by the presence of nucleotide. Signifi- also about 10%. cant shape differences are observed in the presence of ADP-Mg2þor non-hydrolyzable ATP- Mg2þ analog. these methods does not change in the presence of Nucleotide binding disrupts SecA dimer in E. coli lipid vesicles. Flotation analysis confirms lipid vesicles that under these conditions SecA is nonetheless associated stably with the vesicles (and Figure We have used SANS to make direct measure- 1(b)).10 ments of the structure of SecA in vesicles com- The distance distribution function ðPðrÞÞ of apo posed of E. coli phospholipids. In addition to SecA shows that SecA in the presence of lipid providing a model of the physiological membrane, vesicles is a dimer, whereas the dimer dissociates the importance of this lipid composition in promot- upon binding ATP or ADP (Figure 2(b)). The ing ATPase activity27 and translocation28 in vitro asymmetric shape of PðrÞ suggests the SecA dimer has been demonstrated. Furthermore, the depen- is asymmetric. Our SANS results for the SecA dence of the signal peptide-stimulated SecA/lipid dimer without nucleotide in lipid vesicles are thus ATPase activity on a high lipid to protein ratio has in approximate agreement with the SAXS study of been documented27 and is maintained in the SecA in solution,14 although the shape of the SecA studies here. dimer seems to be more expanded in lipid relative The contrast variation experiments require to that in aqueous buffer (160 A˚ versus 150 A˚ ). 2 different volume fractions of H2O to determine Figure 2(a) shows the Guinier plot of SecA at the lipid contrast match point at which the scatter- 2.9 mg/ml of SUV in the presence of non-hydrolyz- ing intensity from the small unilamellar vesicles able ATP analog and Mg2þ. The protein concen- (SUV) is zero. As shown in Figure 1(a), this occurs tration-normalized forward scattering intensity is 2 at 14.3% (v/v) H2O volume fraction. Flotation about half of that of the apo SecA, shown in Figure analysis shows that SecA is lipid-associated (Figure 2(a), indicating that the molecular mass of the ATP 1(b)). Figure 2(a) shows the Guinier plot of 2.9 mg/ analog-bound form of SecA is half of the apo SecA ml of SecA and 23 mg/ml of E. coli SUV in a 14.3% in lipid. The molecular mass calculated from 2 (v/v) H2O buffer solution. For the SecA dimer, the Ið0Þ=cprotein against a molecular mass standard, Guinier analysis was conducted in the region of BSA, shows that the ATP analog-bound form of ˚ 21 Q # 0:027 A corresponding to QRg # 1:2: For SecA is a monomer at the concentration measured 2þ 2þ the ATP-Mg analog and ADP-Mg -bound forms by SANS (Table 1). Similarly, the Ið0Þ=cprotein of of SecA, the Guinier analysis was conducted in SecA in the presence of ADP and Mg2þ is about the region of Q # 0:030 A˚ 21 and Q # 0:034 A˚ 21, half of that of SecA, indicating that the ADP- corresponding to QRg # 1:18 and QRg # 1:08; bound form of SecA is a monomer. In contrast, respectively. SAXS indicated that SecA in aqueous solution The molecular mass standard bovine serum remained dimeric even in the presence of albumin (BSA 68 kDa) was dissolved with 23 mg/ nucleotide.14 2 ml of lipid vesicles in 14.3% H2O (v/v) buffers. The protein concentration-normalized forward The ATP-bound SecA monomer is more scattering intensity ðIð0Þ=cproteinÞ was used to expanded than the ADP-bound SecA monomer calibrate the molecular mass of the different forms of SecA. A calibration of Ið0Þ=cprotein of SecA Based on a Kd of 0.8 mM ADP for SecA in 21 against that of BSA gives a molecular mass of solution, and an estimate of the Kd for the non- 204 kDa for SecA, indicating that it is a dimer hydolyzable ATP analog,29 we have studied SecA at the concentration measured by SANS in the presence of nucleotide sufficient to yield (see Table 1). .97% occupancy of the high-affinity nucleotide- The radius of gyration of SecA in the lipid binding site on SecA. From the Guinier plots vesicles determined by SANS is 45.1(^0.9) A˚ (Figure 2(a)), the radius of gyration of SecA–ATP- 2þ ˚ (Table 1), consistent with the previously reported Mg is Rg ¼ 38:3ð^0:8Þ A, larger than the value of 45.2 A˚ for SecA determined by SAXS radius of gyration of SecA–ADP-Mg2þ of 14 ^ ˚ in aqueous solution. This indicates that the Rg ¼ 31:1ð 0:5Þ A. This suggests that the ATP- radius of gyration of the SecA dimer observed by bound form of the monomeric SecA has a more Nucleotide-induced SecA Conformational Changes 27

expanded global conformation than the ADP-bound, monomeric form. The PðrÞ of the ATP-bound and the ADP-bound forms of SecA show that the SecA dimer dissociates upon binding ATP or ADP (Figure 2(b)), and the maximum dimension of the SecA–ATP-Mg2þ form is larger than that of the SecA–ADP-Mg2þ form (about 125 A˚ and 105 A˚ , respectively). The Kd for the monomer-dimer equilibrium of SecA in aqueous solution in the absence of nucleo- tide is estimated to be in the submicromolar to low micromolar range.13,17 The corresponding values in the presence of lipid vesicles and/or nucleotides remain undefined. A recent study of SecA using fluorescence resonance energy transfer suggests that in the presence of membranes con- taining acidic phospholipids, SecA dissociates into monomers.22 The concentration of SecA used in the fluorescence resonance energy transfer experi- ment was about 1.5 mM, while the concentration used in the SANS experiment was about 28 mM, suggesting that these experiments were run just below and above, respectively, the Kd for dimer dissociation in the absence of nucleotide. Collec- tively, the data suggest that the presence of the lipid environment shifts the Kd upward into the micromolar range and the presence of nucleotide increases the Kd value even further, resulting in a species that is predominantly monomeric. The physiological concentration of SecA is estimated at about 10 mM for the bulk volume of the bacterium16,22 yet is effectively higher at the trans- locon due to the concentrating effect of binding to specific sites comprised of SecYEG. Therefore, our SANS experiments likely provide a good physio- logical representation of the local concentration in this environment. Furthermore, our observation that SecA forms a dimer in the absence of nucleo- tide is consistent with the finding that the dimeric species can associate with the biological membrane in vitro.15 However, since those experiments involved ongoing protein translocation and con- tinual nucleotide turnover, individual SecA states formed during that process could not be distin- guished. An advantage of the strategy we have employed here is the ability to distinguish distinct

by a blue wire model. This blue wire model was calcu- lated from the SANS PðrÞ function using the programs DAMMIN33 and Situs.44 The crystal structure of the apo SecA monomer was docked to the SANS model using the program Situs44 to show that the ATP-bound mono- mer is more expanded than the apo SecA structure. (c) The model of SecA–ADP-Mg2þ calculated from SANS data (blue wire) shows that the ADP-bound monomer is more compact than the apo SecA. These images were made with VMD and are owned by the Theoretical and Computational Biophysics Group, an Figure 3. Models of different forms of SecA. (a) The NIH Resource for Macromolecular Modeling and crystal structure of SecA dimer.18 (b) The SANS-deter- Bioinformatics, at the Beckman Institute, University of mined SecA–ATP-Mg2þ monomer structure represented Illinois at Urbana-Champaign. 28 Nucleotide-induced SecA Conformational Changes structural states under defined conditions. to processively transport successive segments of Moreover, we have been able to characterize the the preprotein. Further, that nucleotide does dimensions of SecA in the presence of a bilayer not induce these conformational changes in the composed of E. coli phospholipids. In this environ- absence of membranes serves to dedicate its func- ment, nucleotide modulates substantial SecA tion in preprotein processivity to the locale of the conformation changes, consistent with previous translocon and inhibit the likelihood that SecA in biochemical analyses in membranes.30 –32 This the cytoplasm participates non-productively in underscores the role of the ATPase activity in the this activity. membrane interactive species and the possibility that the accompanying changes in conformation serve to propel the preprotein through the translocon. The maximum dimension (Dmax) of the SecA– Acknowledgements 2þ ˚ ATP-Mg form is about 125 A and the Dmax of SecA–ADP-Mg2þ form is about 105 A˚ (Figure We thank Sharyn Rusch for helpful discussions 2(b)). We have used the program DAMMIN33 to and critically reading this manuscript. This calculate the envelopes of the different forms of research was supported, in part, by National SecA. Figure 3(a) is the calculated envelope based Institutes of Health grant GM37639 (to D.A.K.). on the crystal structure of B. subtilis SecA in an This work utilized facilities supported, in part, by aqueous environment18 published while we were the National Science Foundation under agreement preparing this manuscript, and Figure 3(b) and (c) no. DMR-9986442. are envelopes of the ATP analog-bound and ADP- bound forms, respectively, of SecA calculated from SANS data. The envelopes shown in Figure 3 References provide a direct view of the structural changes of SecA upon binding to nucleotides. 1. Brundage, L., Hendrick, J. P., Schiebel, E., Driessen, The crystal structure18,19 of SecA can be con- A. J. & Wickner, W. (1990). The purified E. coli sidered to contain three domains: the preprotein- integral membrane protein SecY/E is sufficient for reconstitution of SecA-dependent precursor protein binding domain, the first nucleotide-binding fold translocation. Cell, 62, 649–657. (NBFI), and the second nucleotide-binding fold 2. Brundage, L., Fimmel, C. J., Mizushima, S. & (NBFII). The actual nucleotide-binding site is in Wickner, W. (1992). SecY, SecE, and band 1 the cleft between NBFI and NBFII. In this spirit, form the membrane-embedded domain of Escherichia the structure of SecA resembles the structure of a coli preprotein translocase. J. Biol. Chem. 267, single subunit of GroEL which is composed of 4166–4170. three domains: the apical domain that binds to 3. Douville, K., Price, A., Eichler, J., Economou, A. & unfolded protein, the intermediate domain and Wickner, W. (1995). SecYEG and SecA are the stoi- the equatorial domain. The nucleotide-binding site chiometric components of preprotein translocase. on GroEL is in the cleft formed between the inter- J. Biol. Chem. 270, 20106–20111. 34,35 4. Hanada, M., Nishiyama, K. I., Mizushima, S. & mediate and the equatorial domains. When Tokuda, H. (1994). Reconstitution of an efficient pro- bound to ATP, the intermediate domain of GroEL tein translocation machinery comprising SecA and orients in such a way as to push the apical domain the three membrane proteins, SecY, SecE, and SecG further from the equatorial domain through rigid (p12). J. Biol. Chem. 269, 23625–23631. body movement.34 – 39 In the ADP-bound form, the 5. Akimaru, J., Matsuyama, S., Tokuda, H. & intermediate domain reorients itself to bring the Mizushima, S. (1991). Reconstitution of a protein apical domain and the equatorial domain closer translocation system containing purified SecY, SecE, into a more compact form. Based on our analysis and SecA from Escherichia coli. Proc. Natl Acad. Sci. of the structural analogy of SecA with a single sub- USA, 88, 6545–6549. unit of GroEL, we hypothesize that nucleotide- 6. Cabelli, R. J., Dolan, K. M., Qian, L. P. & Oliver, D. B. (1991). Characterization of membrane-associated and binding induces similar conformation changes in soluble states of SecA protein from wild-type and both proteins and these two proteins may share SecA51(TS) mutant strains of Escherichia coli. J. Biol. common features in the reaction cycles of nucleo- Chem. 266, 24420–24427. tide binding and hydrolysis, and interacting with 7. Mitchell, C. & Oliver, D. (1993). Two distinct ATP- the unfolded proteins. In the case of SecA, these binding domains are needed to promote protein conformational changes are consistent with models export by Escherichia coli SecA ATPase. Mol. Micro- for processive chain movement;40,41 the more biol. 10, 483–497. expanded ATP-bound form is well suited for tra- 8. Ulbrandt, N. D., London, E. & Oliver, D. B. (1992). versing the membrane carrying a portion of the Deep penetration of a portion of Escherichia coli preprotein, while subsequent hydrolysis triggers SecA protein into model membranes is promoted by anionic phospholipids and by partial unfolding. the release of the preprotein, generation of the J. Biol. Chem. 267, 15184–15192. more compact form of SecA and its retraction to 9. Shinkai, A., Mei, L. H., Tokuda, H. & Mizushima, S. the cytoplasmic face of the membrane. ADP-ATP (1991). The conformation of SecA, as revealed by exchange could then serve to initiate another cycle its protease sensitivity, is altered upon interaction of structural changes with insertion and de-insertion with ATP, presecretory proteins, everted membrane Nucleotide-induced SecA Conformational Changes 29

vesicles, and phospholipids. J. Biol. Chem. 266, is a conformationally dynamic and dimeric structure. 5827–5833. EMBO J. 21, 995–1003. 10. Lill, R., Dowhan, W. & Wickner, W. (1990). The 26. Manting, E. H., van der Does, C., Remigy, H., Engel, ATPase activity of SecA is regulated by acidic phos- A. & Driessen, A. J. (2000). SecYEG assembles into a pholipids, SecY, and the leader and mature domains tetramer to form the active protein translocation of precursor proteins. Cell, 60, 271–280. channel. EMBO J. 19, 852–861. 11. Lill, R., Cunningham, K., Brundage, L. A., Ito, K., 27. Miller, A., Wang, L. & Kendall, D. A. (1998). Oliver, D. & Wickner, W. (1989). SecA protein Synthetic signal peptides specifically recognize SecA hydrolyzes ATP and is an essential component of and stimulate ATPase activity in the absence of the protein translocation ATPase of Escherichia coli. preprotein. J. Biol. Chem. 273, 11409–11412. EMBO J. 8, 961–966. 28. Henrick, J. P. & Wickner, W. (1991). SecA protein 12. Cunningham, K. W. & Wickner, W. T. (1989). Specific needs both acidic phospholipids and SecY/E protein recognition of the leader region of precursor proteins for functional high-affinity binding to the Escherichia is required for the activation of translocation ATPase coli plasma membrane. J. Biol. Chem. 266, of Escherichia coli. Proc. Natl Acad. Sci. USA, 86, 24596–24600. 8630–8634. 29. Matsuyama, S., Kimura, E. & Mizushima, S. (1990). 13. Doyle, S. M., Braswell, E. H. & Teschke, C. M. (2000). Complementation of two overlapping fragments of SecA folds via a dimeric intermediate. Biochemistry, SecA, a protein translocation ATPase of Escherichia 39, 11667–11676. coli, allows ATP binding to its amino-terminal region. 14. Shilton, B., Svergun, D. I., Volkov, V. V., Koch, M. H., J. Biol. Chem. 265, 8760–8765. Cusack, S. & Economou, A. (1998). Escherichia coli 30. Ramamurthy, V. & Oliver, D. (1997). Topology of SecA shape and dimensions. FEBS Letters, 436, the integral membrane form of Escherichia coli SecA 277–282. protein reveals multiple periplasmically exposed 15. Dreissen, A. J. M. (1993). SecA, the peripheral sub- regions and modulation by ATP binding. J. Biol. unit of the Escherichia coli precursor protein trans- Chem. 272, 23239–23246. locase, is functional as a dimer. Biochemistry, 32, 31. Eichler, J. & Wickner, W. (1997). Both an N-terminal 13190–13197. 65-kDa domain and a C-terminal 30-kDa domain of 16. Akita, M., Shinkai, A., Matsuyama, S. & Mizushima, SecA cycle into the membrane at SecYEG during S. (1991). SecA, an essential component of the translocation. Proc. Natl Acad. Sci. USA, 94, secretory machinery of Escherichia coli, exists as 5574–5581. homodimer. Biochem. Biophys. Res. Commun. 174, 32. Ahn, T. & Kim, H. (1994). SecA of Escherichia coli 211–216. traverses lipid bilayer of phospholipid vesicles. Bio- 17. Woodbury, R. L., Hardy, S. J. & Randall, L. L. (2002). chem. Biophys. Res. Commun. 203, 326–330. Complex behavior in solution of homodimeric SecA. 33. Svergun, D. I. (1999). Restoring low resolution Protein Sci. 11, 875–882. structure of biological macromolecules from solution 18. Hunt, J. F., Weinkauf, S., Henry, L., Fak, J. J., scattering using simulated annealing. Biophys. J. 76, McNicholas, P., Oliver, D. B. & Deisenhofer, J. 2879–2886. (2002). Nucleotide control of interdomain inter- 34. Braig, K., Otwinowski, Z., Hegde, R., Boisvert, D. C., actions in the conformational reaction cycle of SecA. Joachimiak, A., Horwich, A. L. & Sigler, P. B. (1994). Science, 297, 2018–2026. The crystal structure of the bacterial 19. Sharma, V., Arockiasamy, A., Ronning, D. R., Savva, GroEL at 2.8 A˚ . Nature, 371, 578–586. C. G., Holzenburg, A., Braunstein, M. et al. (2003). 35. Boisvert, D. C., Wang, J., Otwinowski, Z., Horwich, Crystal structure of Mycobacterium tuberculosis SecA, A. L. & Sigler, P. B. (1996). A crystal structure of the a preprotein translocating ATPase. Proc. Natl Acad. bacterial chaperonin GroEL complexed with ATP Sci. USA, 100, 2243–2248. gamma S. Nature Struct. Biol. 3, 170–177. 20. den Blaauwen, T., Fekkes, P., de Wit, J. G., Kuiper, W. 36. Chen, S., Roseman, A. M., Hunter, A. S., Wood, S. P., & Driessen, A. J. M. (1996). Domain interactions of Burston, S. G., Ranson, N. A. et al. (1994). Location the peripheral preprotein translocase subunit SecA. of a folding protein and shape changes in GroEL– Biochemistry, 35, 11994–12004. GroES complexes imaged by cryo-electron 21. Schmidt, M., Ding, H., Ramamurthy, V., Mukerji, I. & microscopy. Nature, 371, 261–264. Oliver, D. (2000). Nucleotide binding activity of 37. Roseman, A. M., Chen, S., White, H., Braig, K. & SecA homodimer is conformationally regulated by Saibil, H. R. (1996). The chaperonin ATPase cycle: temperature and altered by prlD and azi mutations. mechanism of allosteric switching and movements J. Biol. Chem. 275, 15440–15448. of substrate-binding domains in GroEL. Cell, 87, 22. Or, E., Navon, A. & Rapoport, T. (2002). Dissociation 241–251. of the dimeric SecA ATPase during protein trans- 38. Xu, Z., Horwich, A. L. & Sigler, P. B. (1997). The crys- location across the bacterial membrane. EMBO J. 21, tal structure of the asymmetric GroEL-GroES-(ADP)7 4470–4479. chaperonin complex. Nature, 388, 741–750. 23. Benach, J., Chou, Y. T., Fak, J. J., Itkin, A., Nicolae, 39. Ma, J., Sigler, P. B., Xu, Z. & Karplus, M. (2000). A D. D., Smith, P. C. et al. (2003). Phospholipid- dynamic model for the allosteric mechanism of induced monomerization and signal-peptide- GroEL. J. Mol. Biol. 302, 303–313. induced oligomerization of SecA. J. Biol. Chem. 278, 40. Economou, A. & Wickner, W. (1994). SecA promotes 3628–3638. preprotein translocation by undergoing ATP-driven 24. Yahr, T. L. & Wickner, W. T. (2000). Evaluating the cycles of membrane insertion and deinsertion. Cell, oligomeric state of SecYEG in preprotein translocase. 78, 835–843. EMBO J. 19, 4393–4401. 41. Economou, A., Pogliano, J. A., Beckwith, J., Oliver, D. 25. Bessonneau, P., Besson, V., Collinson, I. & Duong, F. & Wickner, W. (1995). SecA membrane cycling at (2002). The SecYEG preprotein translocation channel SecYEG is driven by distinct ATP binding and 30 Nucleotide-induced SecA Conformational Changes

hydrolysis events and is regulated by SecD and SecF. orientation in detergent micelles using small angle Cell, 83, 1171–1181. X-ray scattering. Biophys. J. 77, 1064–1073. 42. Glinka, C., Barker, J., Hammouda, B., Krueger, S., 44. Wriggers, W. & Chaco´n, P. (2001). Using situs for the Moyer, J. & Orts, W. (1998). The 30m SANS instru- registration of protein structures with low-resolution ments at NIST. J. Appl. Crystallog. 31, 430–445. bead models from X-ray solution scattering. J. Appl. 43. Bu, Z. & Engelman, D. M. (1999). A method for Crystallog. 34, 773–776. determining transmembrane helix association and

Edited by I. B. Holland

(Received 11 June 2003; accepted 24 June 2003)