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

Journal logoBIOLOGICAL
CRYSTALLOGRAPHY
ISSN: 1399-0047
Volume 70| Part 8| August 2014| Pages 2197-2203

Pushing the limits of sulfur SAD phasing: de novo structure solution of the N-terminal domain of the ectodomain of HCV E1

CROSSMARK_Color_square_no_text.svg

aDivision of Structural Biology, Wellcome Trust Centre for Human Genetics, University of Oxford, Oxford OX3 7BN, England, and bDiamond Light Source Ltd, Diamond House, Harwell Science and Innovation Campus, Didcot OX11 0DE, England
*Correspondence e-mail: dave@strubi.ox.ac.uk

(Received 23 April 2014; accepted 8 June 2014; online 25 July 2014)

Single-wavelength anomalous dispersion of S atoms (S-SAD) is an elegant phasing method to determine crystal structures that does not require heavy-atom incorporation or selenomethionine derivatization. Nevertheless, this technique has been limited by the paucity of the signal at the usual X-ray wavelengths, requiring very accurate measurement of the anomalous differences. Here, the data collection and structure solution of the N-terminal domain of the ectodomain of HCV E1 from crystals that diffracted very weakly is reported. By combining the data from 32 crystals, it was possible to solve the sulfur substructure and calculate initial maps at 7 Å resolution, and after density modication and phase extension using a higher resolution native data set to 3.5 Å resolution model building was achievable.

1. Introduction

Anomalous dispersion methods are powerful techniques to determine protein structures (Hendrickson, 2013[Hendrickson, W. A. (2013). Acta Cryst. A69, 51-59.]), especially when it is possible to tune the X-ray energy to points close to an absorption edge for atoms within the crystal to maximize the anomalous (Δf′′) and dispersive (Δf′) differences. Multi-wavelength and single-wavelength anomalous dispersion (MAD and SAD) techniques using selenomethionine (SeMet) are nowadays the workhorse methods for the ab initio phasing of macromolecular crystals (Hendrickson et al., 1990[Hendrickson, W. A., Horton, J. R. & LeMaster, D. M. (1990). EMBO J. 9, 1665-1672.]). Despite the success of these methods, some proteins have few or no methionines, or the SeMet-labelled protein may be reluctant to crystallize. In the same manner, selenocysteine-labelled proteins can be expressed in non-auxotrophic Escherichia coli strains (Salgado et al., 2011[Salgado, P. S., Taylor, J. D., Cota, E. & Matthews, S. J. (2011). Acta Cryst. D67, 8-13.]), but this method is likely to encounter the same problems as SeMet-labelled expression, such as lower protein expression, lower solubility or low selenium incorporation in more difficult targets requiring eukaryotic expression systems. Conventional heavy-atom isomorphous replacement methods tend to rely on trial and error (Joyce et al., 2010[Joyce, M. G., Radaev, S. & Sun, P. D. (2010). Acta Cryst. D66, 358-365.]) and usually require testing numerous compounds at different concentrations while keeping the scatterer soluble without damaging the crystals. In contrast, single-wavelength anomalous dispersion of S atoms (S-SAD) does not require the use of selenium-labelled protein or heavy-atom incorporation, as phases can be derived directly from the naturally occurring sulfurs of both cysteines and methionines. Although the S-SAD method was successfully used for the first time more than 30 years ago (Hendrickson & Teeter, 1981[Hendrickson, W. A. & Teeter, M. M. (1981). Nature (London), 290, 107-113.]), the number of de novo structures determined by this method is still limited (Liu et al., 2012[Liu, Q., Dahmane, T., Zhang, Z., Assur, Z., Brasch, J., Shapiro, L., Mancia, F. & Hendrickson, W. A. (2012). Science, 336, 1033-1037.]). The absorption edge of sulfur (∼5 Å) cannot be usefully exploited by conventional synchrotron crystallography beamlines, radiation damage is enhanced at longer wavelengths and absorption becomes severe, so S-SAD is usually carried out at shorter wavelengths (λ = 1.5–2.5 Å). As a direct consequence, the anomalous signal of the sulfur is very weak, so a high signal-to-noise ratio is required for satisfactory measurement of the faint signal. The latter ratio can be improved by increasing the multiplicity; however, poorly diffracting crystals require greater X-ray doses and thus obtaining high-multiplicity data sets is often not possible from a single crystal. In order to overcome this problem, the anomalous differences can be recorded from multiple isomorphous crystals until the desired multiplicity is reached while keeping the radiation damage low (Liu et al., 2012[Liu, Q., Dahmane, T., Zhang, Z., Assur, Z., Brasch, J., Shapiro, L., Mancia, F. & Hendrickson, W. A. (2012). Science, 336, 1033-1037.], 2013[Liu, Q., Liu, Q. & Hendrickson, W. A. (2013). Acta Cryst. D69, 1314-1332.]). A second method to enhance the anomalous differences in the face of radiation damage is to use the inverse-beam data-collection strategy (Hendrickson et al., 1989[Hendrickson, W. A., Pähler, A., Smith, J. L., Satow, Y., Merritt, E. A. & Phizackerley, R. P. (1989). Proc. Natl Acad. Sci. USA, 86, 2190-2194.]). The Friedel mates (h, k, l) and (−h, −k, −l) are recorded in small wedges at φ and φ + 180°, ensuring that Friedel pairs are recorded close in time while minimizing the difference in absorption effects and radiation damage.

S-SAD phasing was applied to determine the structure of the N-terminal domain of the ectodomain of Hepatitis C virus envelope glycoprotein E1 (HCV nE1). The HCV envelope glycoproteins E1 and E2 are located on the surface of the virions and are responsible for binding of the virus to the host cells and membrane fusion. Although HCV is a major global health problem, its mechanism of fusion is still not known owing to the lack of structural knowledge of these two glycoproteins. HCV nE1 is composed of 79 residues and contains no methionines, which makes this construct unsuitable for SeMet phasing. Heavy-atom soaking experiments were attempted but failed to show any useful anomalous signal for substructure determination; therefore, efforts were focused on S-SAD methods.

2. Methods and results

2.1. Cloning, expression and protein purification

DNA coding for the ectodomain of HCV E1 (residues 1–79) was synthesized with a mutation at one of the glycosylation sites (N43Q) and was cloned into the pHLsec vector (Aricescu et al., 2006[Aricescu, A. R., Lu, W. & Jones, E. Y. (2006). Acta Cryst. D62, 1243-1250.]). The construct containing a C-terminal His6 tag (Fig. 1[link]a) was transiently expressed in HEK293T cells in the presence of 5 µM kifunensine to limit N-glycosylation of the remaining sites (Toronto Research Chemicals, North York, Ontario, Canada). Ni2+-affinity purification (FF Chelating Sepharose resin, GE Healthcare) was followed by TEV protease and endoglycosidase F1 treatment before size-exclusion chromatography on a Superdex 75 column (GE Healthcare). The protein was estimated to be greater than 95% pure by SDS–PAGE (Fig. 1[link]b). 3-(1-Pyridino)-1-propanesulfonate (NDSB 201; Soltec Ventures Inc.) was added to HCV nE1 to a final concentration of 300 mM in order to reach concentrations of between 17 and 22 mg ml−1.

[Figure 1]
Figure 1
Construct and data-collection details. (a) Amino-acid sequence of HCV nE1. Cysteines, glycosylation sites and the N43Q mutation are shown in green, blue and red, respectively. The extra residues resulting from cloning are coloured light blue (TEV cleavage site) and pink. (b) 15% reducing (R) and nonreducing (NR) SDS–PAGE gels showing the purity of deglycosylated HCV nE1. (c) Typical HCV nE1 crystal. The red ellipse represents the size of the beam. (d) DISTL plot showing the number of spots and estimated resolution for each image (or ω) in a representative wedge (Zhang et al., 2006[Zhang, Z., Sauter, N. K., van den Bedem, H., Snell, G. & Deacon, A. M. (2006). J. Appl. Cryst. 39, 112-119.]). The number of found spots (red), potential Bragg candidates (green) and the resolution (blue) are depicted as crosses.

2.2. Crystallization

A Cartesian Technologies MIC4000 robot was used to set up high-throughput crystallization trials using the sitting-drop vapour-diffusion method at 294 K in 96-well plates (Greiner Bio-One Ltd, Stonehouse, England; Walter et al., 2003[Walter, T. S., Diprose, J., Brown, J., Pickford, M., Owens, R. J., Stuart, D. I. & Harlos, K. (2003). J. Appl. Cryst. 36, 308-314.], 2005[Walter, T. S. et al. (2005). Acta Cryst. D61, 651-657.]). Initial crystal hits for HCV nE1 N43Q were obtained in 15%(w/v) PEG 1500, 3.6%(w/v) PEG 4000, 0.05 M sodium acetate pH 4.8 (Pi–PEG screen, Jena Bioscience). Crystals of hexagonal morphology appeared after a few days but diffracted extremely weakly and appeared to be twinned. The same condition after some two weeks gave crystals of tetragonal morphology, which were optimized using an additive screen (Hampton Research). Addition of 100 nl of 6–8% 2,5-hexanediol or 1,6-hexanediol to the initial condition improved the size of the crystals to 110 × 30 × 10 µm (Fig. 1[link]c). Crystals were flash-cooled in liquid nitrogen using 25%(v/v) ethylene glycol in the reservoir solution as a cryoprotectant.

2.3. Data collection

An initial data set was recorded at 100 K on the I24 beamline at Diamond Light Source (DLS), Didcot, England at a wavelength of 0.9796 Å using a PILATUS 6M detector (DECTRIS) with the crystal-to-detector distance set to 623.5 mm to cover diffraction to 3 Å resolution at the detector edge. A total crystal rotation range of 90° was collected from a single crystal with an exposure time of 0.2 s per 0.1° (100% beam transmission: 1012 photons s−1 with a beam size of 30 × 30 µm). The space group P41212 (or P43212) and unit-cell parameters a = b = 105.0, c = 204.8 Å, α = β = γ = 90° were obtained by processing the data with HKL-2000 (Otwinowski & Minor, 1996[Otwinowski, Z. & Minor, W. (1996). Methods Enzymol. 276, 307-326.]). The data extended to ∼3.5 Å resolution (Table 1[link]).

Table 1
Data-collection statistics

Values in parentheses are for the highest resolution shell.

  S-SAD S-SAD Native
Data collection
 Beamline I04, DLS I04, DLS I24, DLS
 Space group P41212 P41212 P41212
 Unit-cell parameters (Å, °) a = b = 105.2, c = 204.4, α = β = γ = 90 a = b = 105.5, c = 204.8, α = β = γ = 90 a = b = 105.0, c = 204.7, α = β = γ = 90
 No. of crystals 1 [1-wedge series] 32 [64-wedge series] 1
 Wavelength (Å) 1.7712 1.7712 0.9686
 Resolution (Å) 42.7–4.5 (4.64–4.52) 60.3–4.2 (4.32–4.21) 50–3.5 (3.63–3.50)
 No. of unique reflections 10723 (625) 15823 (1033) 15137 (1471)
 Completeness (%) 92.8 (82.2) 99.4 (96.3) 99.5 (99.5)
 Multiplicity 3.1 (2.4) 121.5 (4.2) 6.2 (6.2)
 〈I/σ(I)〉 6.8 (2.5) 33.3 (3.6) 17.6 (2.2)
Rmerge (%) 9.8 (24.1) 16.0 (35.2) 13.5 (81.0)
Rp.i.m. (%) 7.0 (20.9) 1.7 (24.0) 5.5 (35.2)
 CC1/2, highest resolution shell 0.90 0.82 0.66
Refinement
 Resolution (Å)     31.3–3.5
Rwork/Rfree (%)     21.6/23.7
 R.m.s.d., bond lengths (Å)     0.008
 R.m.s.d., angles (°)     1.13
 Mean B factor (Å2)     88.4
 Wilson B factor (Å2)     118.2
 Ramachandran plot (%)
  Favoured     97.4
  Allowed     100
  Outliers     0
Rmerge = [\textstyle \sum_{hkl}\sum_{i}|I_{i}(hkl)- \langle I(hkl)\rangle|/][\textstyle \sum_{hkl}\sum_{i}I_{i}(hkl)], where Ii(hkl) is the ith measurement of reflection hkl and 〈I(hkl)〉 is the weighted average of all measured reflections.

HCV nE1 contains 79 residues, two glycosylation sites, four cysteines and no methionines (Fig. 1[link]a). For a solvent content of 52%, the asymmetric unit would comprise 13 molecules (VM of 2.4 Å3 Da−1), although the very weak diffraction suggested that the solvent content might be higher. From comparison of reducing and nonreducing SDS–PAGE gels (Fig. 1[link]b), HCV nE1 forms covalent dimers (in agreement with size-exclusion chromatography; data not shown). We did not know whether all of the cysteine residues would be involved in disulfide bonds, but speculated that this was quite likely and recognized that this would enhance the phasing power at very low resolution, where the bonded atoms would scatter coherently, and simplify the determination of the sulfur sub­structure. A calculated Bijvoet ratio of 1.1% (for four free cysteines, or 1.7% for four cysteines involved in disulfide bridges) for the total reflection intensities led us to target an overall signal-to-noise ratio of at least 30 for effective phasing (this guide figure was based on the expectation that the substructure could be determined from the stronger lower resolution reflections). For S-SAD experiments, data sets from 32 randomly orientated crystals were recorded at a wavelength of 1.7712 Å using the inverse-beam method on the I04 beamline at DLS using a PILATUS 6M detector (DECTRIS) with the crystal-to-detector distance set to 560 mm to cover diffraction to 4.5 Å resolution at the detector edge (a helium path was not used). A beam size of 80 × 45 µm was used with a flux of 1.5–2.0 × 1011 photons s−1. Each crystal was rotated 180° from the initial position every 5° to measure Friedel pairs. On average a total of 90° was collected per crystal in two wedge series (A and B) of 9 × 5° each with a rotation of 0.05° and an exposure time of 0.05 s per frame. The 64-wedge series (57 600 frames in total) was auto-processed and merged with xia2 (Winter et al., 2013[Winter, G., Lobley, C. M. C. & Prince, S. M. (2013). Acta Cryst. D69, 1260-1273.]) with good statistics: overall Rmerge, completeness and multiplicity of 0.16, 0.99 and 121, respectively. The quality of the merging was reflected in the small number of rejections (0.25%). Data-collection details are shown in Table 1[link], which also reports, for comparison purposes, statistics for a typical S-SAD wedge. The rationale for the choice of data-collection parameters is given below.

In order to mitigate absorption effects at longer wavelengths while being able to collect a useful sulfur anomalous signal, the beam wavelength was tuned to 1.77 Å (f′′ = 0.7 electrons). It was also crucial to know the lifetime of the crystals when exposed to X-rays. At the selenium edge wavelength at I24, HCV nE1 crystals lasted about 180 s, but to test the behaviour of the crystals at λ = 1.77 Å at I04 we assessed the crystal decay by looking at the number of observed spots per image and finally collected 90 s per crystal (Fig. 1[link]d). With the aim of maximizing the signal-to-noise ratio, very small rotation angles of 0.05° per image were collected on a PILATUS 6M detector (DECTRIS) operating in shutterless mode (across the 100 images of each 5° wedge) and in order to collect the data sets quickly we used a non-attenuated beam (1.5–2.0 × 1011 photons s−1) with a very limited exposure time of 0.05 s. A beam size of 80 × 50 µm was used to match the size of the crystals.

Because it was not possible to obtain high multiplicity from a single HCV nE1 crystal, an overall multiplicity of 121 (4.2 in the outer shell) was built up by collecting data sets from 32 crystals. The scaling of all data sets was of excellent quality, with Rmerge and Rp.i.m. values of 0.16 and 0.017, respectively, for the overall data and of 0.35 and 0.24, respectively, for the outer shell. Although the crystal-to-detector distance was set to record reflections to 4.5 Å resolution at the edge, multiple crystals in random orientations permitted full coverage of reciprocal space and allowed the resolution to be extended to 4.2 Å (the corner of the detector) with a CC1/2 (Karplus & Diederichs, 2012[Karplus, P. A. & Diederichs, K. (2012). Science, 336, 1030-1033.]) of 0.99 overall and of 0.82 for the highest resolution shell. The anomalous signal extends to 6.7 Å resolution according to XSCALE (Kabsch, 2010a[Kabsch, W. (2010a). Acta Cryst. D66, 125-132.],b[Kabsch, W. (2010b). Acta Cryst. D66, 133-144.]) {[|F(+) − F(−)|/σ] of 1.1 with an anomalous correlation of 31%}, with an overall anomalous multiplicity of 66. Combining multiple crystals for low-resolution phasing has previously been shown to be useful for structure determination in difficult cases (Liu et al., 2013[Karplus, P. A. & Diederichs, K. (2012). Science, 336, 1030-1033.]). An efficient inverse-beam mode method was specifically implemented at the beamline for automatic data collection which allows the recording of accurate Friedel pairs to be prioritized over data completeness. Each crystal was rotated 180° from the starting position every 5° and a total of 90° was collected per crystal in two wedges of 45°. It was essential that the crystals were isomorphous in order to merge them; indeed, merging data from sufficiently non-isomorphous crystals would degrade the anomalous signal. Programs such as BLEND (Foadi et al., 2013[Foadi, J., Aller, P., Alguel, Y., Cameron, A., Axford, D., Owen, R. L., Armour, W., Waterman, D. G., Iwata, S. & Evans, G. (2013). Acta Cryst. D69, 1617-1632.]) select the optimal clusters of data sets from multiple crystals prior to scaling and merging. In our case, the 32 crystals (64 sweeps) were analysed for isomorphism, and all wedges shared, on pairwise comparison, correlation coefficients of at least 0.92 (0.97 on average) and r.m.s. deviations of 0.26 and 0.45 Å in the a and c unit-cell parameters, respectively. BLEND calculated a linear cell variation of 1.25% between the 64 sweeps (this is the maximum linear change in the diagonals on the three independent cell faces; Foadi et al., 2013[Foadi, J., Aller, P., Alguel, Y., Cameron, A., Axford, D., Owen, R. L., Armour, W., Waterman, D. G., Iwata, S. & Evans, G. (2013). Acta Cryst. D69, 1617-1632.]), suggesting that all 64 wedges should be merged in xia2 (Winter et al., 2013[Winter, G., Lobley, C. M. C. & Prince, S. M. (2013). Acta Cryst. D69, 1260-1273.]) to give the statistics shown in Table 1[link].

2.4. Structure determination and refinement

The sulfur substructure was determined using the HKL2MAP graphical interface (Pape & Schneider, 2004[Pape, T. & Schneider, T. R. (2004). J. Appl. Cryst. 37, 843-844.]) with SHELXC, SHELXD and SHELXE (Sheldrick, 2010[Sheldrick, G. M. (2010). Acta Cryst. D66, 479-485.]). SHELXC showed a weak anomalous signal extending to about 6.5–7 Å resolution (Fig. 2[link]a). It was initially difficult to locate any sulfur sites with SHELXD as the crystals have an even higher solvent content than expected (six molecules in the asymmetric unit, corresponding to 75% solvent content with a VM of 4.9 Å3 Da−1); thus, the number of sites searched for was initially overestimated. After performing multiple runs (1000 trials per run) with different numbers of heavy-atom sites and resolution cutoffs, a solution could be obtained for 12 S atoms at 7 Å resolution (in the most favourable case the success rate was 0.8%; Fig. 2[link]b). The main criterion for selecting a probable number of sulfur sites in the asymmetric unit was to select the SHELXD runs which gave the highest CCall and CCweak and to judge the number of sites by the occupancies. For six molecules in the asymmetric unit, we expected that the 24 sulfurs might be involved in disulfide bonding, but at such low resolution a disulfide bond would scatter coherently as a single heavy atom (Debreczeni et al., 2003[Debreczeni, J. É., Bunkóczi, G., Ma, Q., Blaser, H. & Sheldrick, G. M. (2003). Acta Cryst. D59, 688-696.]; Usón et al., 2003[Usón, I., Schmidt, B., von Bülow, R., Grimme, S., von Figura, K., Dauter, M., Rajashankar, K. R., Dauter, Z. & Sheldrick, G. M. (2003). Acta Cryst. D59, 57-66.]; the transverse coherence length of the X-ray beam is more than four orders of magnitude greater than this bond length). The correctness of the solution was confirmed by SHELXE, which showed a separation in the map contrast between the two hands (0.377 versus 0.290), implying that the correct space group was P41212 and not P43212 (Fig. 2[link]c); nevertheless, the initial maps were not readily interpretable (Fig. 2[link]d).

[Figure 2]
Figure 2
HKL2MAP profiles. (a) d′′/sig(d′′) as a function of resolution. The graph shows the signal to noise from the anomalous differences. In the red part of the graph the anomalous signal is considered to be nonexistent. (b) Profiles of correlation coefficients between observed and calculated Bijvoet differences. (c) Contrast between the variance in the electron density in the protein region and in the solvent region for a given phase set as a function of cycle number with phases calculated based on the original (red) or inverted (blue) substructure. (d) Initial experimental electron-density maps at 7 Å resolution (original) contoured at at 1σ obtained from SHELXE; the final model has been displayed to assess the map quality. (e) d′′/sig(d′′) as a function of resolution as in (a) but using calculated anomalous differences from the final refined HCV nE1 model.

SAD phasing was performed by phenix.autosol (Adams et al., 2002[Adams, P. D., Grosse-Kunstleve, R. W., Hung, L.-W., Ioerger, T. R., McCoy, A. J., Moriarty, N. W., Read, R. J., Sacchettini, J. C., Sauter, N. K. & Terwilliger, T. C. (2002). Acta Cryst. D58, 1948-1954.]) using the sulfur sites obtained by HKL2MAP (Pape & Schneider, 2004[Pape, T. & Schneider, T. R. (2004). J. Appl. Cryst. 37, 843-844.]). It was essential to cut the resolution to 7 Å and set the solvent content to 0.7 to obtain initial phases (Fig. 3[link]a) and only then extend to the full resolution (4.2 Å); however, the software was not able to automatically determine the NCS operators, so rebuilding was not feasible. Nonetheless, it was possible to identify density possibly corresponding to α-helices. Six α-helices were located in the map and manually fitted using Coot (Emsley & Cowtan, 2004[Emsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126-2132.]), keeping the same orientation within each monomer (at this resolution the helix directionality could not be determined); noncrystallographic symmetry (NCS) operators were then calculated using phenix.find_ncs_operators (Adams et al., 2002[Adams, P. D., Grosse-Kunstleve, R. W., Hung, L.-W., Ioerger, T. R., McCoy, A. J., Moriarty, N. W., Read, R. J., Sacchettini, J. C., Sauter, N. K. & Terwilliger, T. C. (2002). Acta Cryst. D58, 1948-1954.]). These were then input to phenix.autobuild (Adams et al., 2002[Adams, P. D., Grosse-Kunstleve, R. W., Hung, L.-W., Ioerger, T. R., McCoy, A. J., Moriarty, N. W., Read, R. J., Sacchettini, J. C., Sauter, N. K. & Terwilliger, T. C. (2002). Acta Cryst. D58, 1948-1954.]) with the higher resolution data set (FP and SIGFP), initial maps (phases) and heavy-atom positions (which helped with the NCS determination). Density modification using a solvent content of 75%, sixfold NCS averaging and extension of the resolution to that of the native data set (3.5 Å resolution) resulted in interpretable maps (Fig. 3[link]b). Secondary structures were clearly visible (Fig. 3[link]b) and a partial structure could be built using Coot (Emsley & Cowtan, 2004[Emsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126-2132.]). Refinement using autoBUSTER with local structure symmetry and external (S-SAD) phase restraints (Bricogne et al., 2008[Bricogne, G., Blanc, E., Brandl, M., Flensburg, C., Keller, P., Paciorek, W., Roversi, P., Smart, O. S., Vonrhein, C. & Womack, T. (2008). BUSTER. Cambridge: Global Phasing Ltd.]), alternating with rebuilding using Coot, taking into account cysteine positions (four per monomer, all involved in disulfide bonds) and glycan positions (two per monomer), led to a reliable structure and excellent quality electron-density maps. Refinement statistics are given in Table 1[link]. As expected, the quality of the maps benefited from the 75% solvent content (Watanabe et al., 2005[Watanabe, N., Kitago, Y., Tanaka, I., Wang, J., Gu, Y., Zheng, C. & Fan, H. (2005). Acta Cryst. D61, 1533-1540.]) and sixfold NCS (Figs. 3[link]c and 3[link]d). The structure will be described elsewhere (manuscript submitted) and the coordinates and structure factors have been deposited in the Protein Data Bank as entry 4uoi .

[Figure 3]
Figure 3
Improvement of electron-density maps. The blue meshes show the electron density contoured at 1σ. (a) Electron-density maps at 7 Å resolution after density modification by phenix.autosol using a solvent content of 75%. (b) Electron-density maps at 3.5 Å resolution after density modification by phenix.autobuild using sixfold NCS. (c) Final 2|Fo| − |Fc| electron-density maps at 3.5 Å resolution after refinement with autoBUSTER. (d) Structure of HCV nE1 fitted into the electron-density maps described in (c). The six monomers composing the aymmetric unit are coloured differently.

From the refined structure, we calculated theoretical anomalous differences using phenix.fmodel (Adams et al., 2002[Adams, P. D., Grosse-Kunstleve, R. W., Hung, L.-W., Ioerger, T. R., McCoy, A. J., Moriarty, N. W., Read, R. J., Sacchettini, J. C., Sauter, N. K. & Terwilliger, T. C. (2002). Acta Cryst. D58, 1948-1954.]) in order to plot the calculated anomalous signal against resolution. The structure factors were also calculated from structures in which the disulfide bonds were disrupted by rotating each side chain by 180° or by placing S atoms 10 Å away from each other (Fig. 4[link]). This shows the expected marked increase in anomalous signal at low resolution (below 5.5 Å) when the sulfurs are involved in disulfide bonding, reflecting the coherent diffraction of two sulfurs. At higher resolution this coherence is lost.

[Figure 4]
Figure 4
Calculated anomalous differences. Calculated d′′/sig(d′′) from refined structures as a function of resolution. The graph shows the signal to noise from the anomalous differences. In the red part of the graph the anomalous signal is considered to be nonexistent. The d′′/sig(d′′) calculated from the final structure, from a structure with cysteine side chains flipped by 180° and from a structure with S atoms from disulfide bonds moved 10 Å away from each other are coloured blue, green and red, respectively.

3. Conclusions

Recent developments in synchrotron instrumentation and crystallographic software have helped to improve the sulfur SAD phasing method, which is in principle the best technique for structure solution as most native crystals can be directly used for phasing. Practically, the approach is limited by a number of different factors. The work reported here shows that useful phasing can be obtained without the need for high-resolution diffraction, or indeed strongly diffracting crystals, if careful data collection is carried out in order to obtain a highly redundant data set from mutiple crystals; indeed, the useful anomalous signal of HCV nE1 crystals did not extend to better than 6.5 Å resolution. The nature of the crystals is also very important; in our case we benefitted from isomorphous crystals, facilitating the scaling and merging of the data, whilst a high solvent content and NCS improved the quality of the early maps. We expect that future hardware and software development will increase the success rate of sulfur phasing and increasingly render it the method of choice for ab initio phasing.

Supporting information


Footnotes

Current address: EMBL Grenoble, 71 avenue des Martyrs, 38042 Grenoble, France.

Acknowledgements

Geoff Sutton and Tom Walter are thanked for valuable technical assistance. We thank the staff of beamlines I04 and I24 at the Diamond Light Source synchrotron for technical support. This work was supported by the Medical Research Council (MRC; grant G1000099), and the Wellcome Trust provided administrative support (grant 075491/Z/04).

References

First citationAdams, P. D., Grosse-Kunstleve, R. W., Hung, L.-W., Ioerger, T. R., McCoy, A. J., Moriarty, N. W., Read, R. J., Sacchettini, J. C., Sauter, N. K. & Terwilliger, T. C. (2002). Acta Cryst. D58, 1948–1954.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationAricescu, A. R., Lu, W. & Jones, E. Y. (2006). Acta Cryst. D62, 1243–1250.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBricogne, G., Blanc, E., Brandl, M., Flensburg, C., Keller, P., Paciorek, W., Roversi, P., Smart, O. S., Vonrhein, C. & Womack, T. (2008). BUSTER. Cambridge: Global Phasing Ltd.  Google Scholar
First citationDebreczeni, J. É., Bunkóczi, G., Ma, Q., Blaser, H. & Sheldrick, G. M. (2003). Acta Cryst. D59, 688–696.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEmsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126–2132.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFoadi, J., Aller, P., Alguel, Y., Cameron, A., Axford, D., Owen, R. L., Armour, W., Waterman, D. G., Iwata, S. & Evans, G. (2013). Acta Cryst. D69, 1617–1632.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationHendrickson, W. A. (2013). Acta Cryst. A69, 51–59.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationHendrickson, W. A., Horton, J. R. & LeMaster, D. M. (1990). EMBO J. 9, 1665–1672.  CAS PubMed Web of Science Google Scholar
First citationHendrickson, W. A., Pähler, A., Smith, J. L., Satow, Y., Merritt, E. A. & Phizackerley, R. P. (1989). Proc. Natl Acad. Sci. USA, 86, 2190–2194.  CrossRef CAS PubMed Web of Science Google Scholar
First citationHendrickson, W. A. & Teeter, M. M. (1981). Nature (London), 290, 107–113.  CrossRef CAS Web of Science Google Scholar
First citationJoyce, M. G., Radaev, S. & Sun, P. D. (2010). Acta Cryst. D66, 358–365.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKabsch, W. (2010a). Acta Cryst. D66, 125–132.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKabsch, W. (2010b). Acta Cryst. D66, 133–144.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKarplus, P. A. & Diederichs, K. (2012). Science, 336, 1030–1033.  Web of Science CrossRef CAS PubMed Google Scholar
First citationLiu, Q., Dahmane, T., Zhang, Z., Assur, Z., Brasch, J., Shapiro, L., Mancia, F. & Hendrickson, W. A. (2012). Science, 336, 1033–1037.  Web of Science CrossRef CAS PubMed Google Scholar
First citationLiu, Q., Liu, Q. & Hendrickson, W. A. (2013). Acta Cryst. D69, 1314–1332.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationOtwinowski, Z. & Minor, W. (1996). Methods Enzymol. 276, 307–326.  CrossRef Web of Science Google Scholar
First citationPape, T. & Schneider, T. R. (2004). J. Appl. Cryst. 37, 843–844.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSalgado, P. S., Taylor, J. D., Cota, E. & Matthews, S. J. (2011). Acta Cryst. D67, 8–13.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSheldrick, G. M. (2010). Acta Cryst. D66, 479–485.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationUsón, I., Schmidt, B., von Bülow, R., Grimme, S., von Figura, K., Dauter, M., Rajashankar, K. R., Dauter, Z. & Sheldrick, G. M. (2003). Acta Cryst. D59, 57–66.  Web of Science CrossRef IUCr Journals Google Scholar
First citationWalter, T. S., Diprose, J., Brown, J., Pickford, M., Owens, R. J., Stuart, D. I. & Harlos, K. (2003). J. Appl. Cryst. 36, 308–314.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWalter, T. S. et al. (2005). Acta Cryst. D61, 651–657.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWatanabe, N., Kitago, Y., Tanaka, I., Wang, J., Gu, Y., Zheng, C. & Fan, H. (2005). Acta Cryst. D61, 1533–1540.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWinter, G., Lobley, C. M. C. & Prince, S. M. (2013). Acta Cryst. D69, 1260–1273.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationZhang, Z., Sauter, N. K., van den Bedem, H., Snell, G. & Deacon, A. M. (2006). J. Appl. Cryst. 39, 112–119.  Web of Science CrossRef CAS IUCr Journals Google Scholar

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

Journal logoBIOLOGICAL
CRYSTALLOGRAPHY
ISSN: 1399-0047
Volume 70| Part 8| August 2014| Pages 2197-2203
Follow Acta Cryst. D
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds