crystallization communications\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL BIOLOGY
COMMUNICATIONS
ISSN: 2053-230X

Expression, purification, crystallization and preliminary crystallographic analysis of BipD, a component of the Burkholderia pseudomallei type III secretion system

CROSSMARK_Color_square_no_text.svg

aLaboratory of Molecular Biophysics, Department of Biochemistry, University of Oxford, South Parks Road, Oxford OX1 3QU, England, bDivision of Microbiology, Institute for Animal Health, Compton Laboratory, Berkshire RG20 7NN, England, and cSir William Dunn School of Pathology, University of Oxford, South Parks Road, Oxford OX1 3RE, England
*Correspondence e-mail: susan.lea@biop.ox.ac.uk

(Received 28 April 2006; accepted 12 July 2006; online 11 August 2006)

A construct consisting of residues 10–310 of BipD, a component of the Burkholderia pseudomallei type III secretion system (T3SS), has been overexpressed as a GST fusion, cleaved from the GST tag and purified. Crystals were grown of native and selenomethionine-labelled BipD. The crystals grow in two different polymorphs from the same condition. The first polymorph belongs to space group C222, with unit-cell parameters a = 103.98, b = 122.79, c = 49.17 Å, a calculated Matthews coefficient of 2.4 Å3 Da−1 (47% solvent content) and one molecule per asymmetric unit. The second polymorph belongs to space group P21212, with unit-cell parameters a = 136.47, b = 89.84, c = 50.15 Å, and a calculated Matthews coefficient of 2.3 Å3 Da−1 (45% solvent content) for two molecules per asymmetric unit (analysis of the self-rotation function indicates the presence of a weak twofold non-crystallographic symmetry axis in this P21212 form). The native crystals of both forms give diffraction data to 2.7 Å resolution, while the SeMet-labelled P21212 crystals diffract to 3.3 Å resolution. A K2PtCl4 derivative of the P21212 form was also obtained and data were collected to 2.7 Å with radiation of wavelength λ = 0.933 Å. The Pt-derivative anomalous difference Patterson map revealed two self-peaks on the Harker sections.

1. Introduction

Type III secretion systems (T3SSs) are essential virulence determinants of many Gram-negative bacterial pathogens. The T3SS is required to translocate virulence effectors into the host cell. The T3SS consists of a `needle complex' composed of an external hollow needle held within a basal body that traverses both bacterial membranes. Secretion is activated by contact of the tip of the needle with host cells, resulting in the formation of a pore in the host-cell membrane that is contiguous with the needle. Other effector proteins are injected via this apparatus directly into the host-cell cytoplasm (for a review, see Johnson et al., 2005[Johnson, S., Deane, J. E. & Lea, S. M. (2005). Curr. Opin. Struct. Biol. 15, 700-­707.]).

Among Gram-negative bacteria that possess a T3SS, the Burkholderia species are highly pathogenic to humans. They have been listed as biological risk category B agents and as a consequence of their infectivity by the respiratory route are considered potential bioterror agents (Rotz et al., 2002[Rotz, L. D., Khan, A. S., Lillibridge, S. R., Ostroff, S. M. & Hughes, J. M. (2002). Emerg. Infect. Dis. 8, 225-230.]). B. pseudomallei causes melioid­osis in humans, a disease endemic in Southeast Asia and Northern Australia. This disease presents in a variety of ways from subacute and chronic suppurative infections to a rapidly fatal septicaemia (White, 2003[White, N. J. (2003). Lancet, 361, 1715-1722.]). B. mallei causes the zoonotic disease glanders, mainly in horses. B. mallei can also infect humans, an infection that is almost invariably fatal if untreated (Wilkinson, 1981[Wilkinson, L. (1981). Med. Hist. 25, 363-384.]).

The B. pseudomallei needle is composed of the ∼9 kDa protein BsaL, the structure of which has been determined by NMR (Zhang et al., 2006[Zhang, L., Wang, Y., Picking, W. L., Picking, W. D. & De Guzman, R. N. (2006). J. Mol. Biol. 359, 322-330.]) and is homologous both in sequence and structure to the Shigella flexneri needle component MxiH. We have recently determined the crystal structure of MxiH and assembled a molecular model of the T3SS needle, docking its crystal structure into a 16 Å EM reconstruction of the S. flexneri needle (Cordes et al., 2005[Cordes, F. S., Daniell, S., Kenjale, R., Saurya, S., Picking, W. L., Picking, W. D., Booy, F., Lea, S. M. & Blocker, A. (2005). J. Mol. Biol. 354, 206-211.]; Deane, Cordes et al., 2006[Deane, J. E., Cordes, F. S., Roversi, P., Johnson, S., Kenjale, R., Picking, W. D., Picking, W. L., Lea, S. M. & Blocker, A. (2006). Acta Cryst. F62, 302-305.]; Deane, Roversi et al., 2006[Deane, J. E., Roversi, P., Cordes, F. S., Johnson, S., Kenjale, R., Daniell, S., Booy, F., Picking, W. D., Picking, W. L., Blocker, A. J. & Lea, S. M. (2006). Proc. Natl Acad. Sci. USA, doi:10.1073/pnas.0602689103.]). As part of our ongoing structural effort to further characterize the T3SS needle machinery, we have now endeavoured to express, purify and crystallize the B. pseudomallei BipD protein, a key component of the translocation apparatus in this bacterium. BipD is required for full virulence in murine models of melioidosis and mutation of the B. pseudomallei bipD gene impairs invasion of epithelial cells in vitro (Stevens & Galyov, 2004[Stevens, M. P. & Galyov, E. E. (2004). Int. J. Med. Microbiol. 293, 549-555.]; Stevens et al., 2004[Stevens, M. P., Haque, A., Atkins, T., Hill, J., Wood, M. W., Easton, A., Nelson, M., Underwood-Fowler, C., Titball, R. W., Bancroft, G. J. & Galyov, E. E. (2004). Microbiology, 150, 2669-2676.]).

There are no close BipD sequence homologues of known structure, the closest (13% sequence identity) being LcrV (Derewenda et al., 2004[Derewenda, U., Mateja, A., Devedjiev, Y., Routzahn, K. M., Evdokimov, A. G., Derewenda, Z. S. & Waugh, D. S. (2004). Structure, 12, 301-306.]), a T3SS component of Yersinia pestis, which has been shown to be located at the tip of the T3SS needle and is similarly needed for cell invasion (Mota, 2006[Mota, L. J. (2006). Trends Microbiol. 14, 197-200.]). The closest sequence homologues of BipD are Shigella IpaD and Salmonella SipD (36% and 33% sequence identity to BipD, respectively); together with some enteropathogenic Escherichia coli and Chromobacterium violaceum proteins, they form the T3SS protein family called `invasion plasmid antigen IpaD proteins' (Bateman et al., 2004[Bateman, A., Coin, L., Durbin, R., Finn, R. D., Hollich, V., Griffiths-Jones, S., Khanna, A., Marshall, M., Moxon, S., Sonnhammer, E. L., Studholme, D. J., Yeats, C. & Eddy, S. R. (2004). Nucleic Acids Res. 32, D138-D141.]). A three-dimensional structure of Burkholderia BipD, together with the structure of Yersinia LcrV (Derewenda et al., 2004[Derewenda, U., Mateja, A., Devedjiev, Y., Routzahn, K. M., Evdokimov, A. G., Derewenda, Z. S. & Waugh, D. S. (2004). Structure, 12, 301-306.]), should increase our understanding of their role in T3SS pathogenesis.

2. Experimental procedures

2.1. Protein expression and purification

The construction of the pGEXBipD plasmid used for expression of the GST-BipD fusion protein has been described previously (Stevens et al., 2002[Stevens, M. P., Wood, M. W., Taylor, L. A., Monaghan, P., Hawes, P., Jones, P. W., Wallis, T. S. & Galyov, E. E. (2002). Mol. Microbiol. 46, 649-659.]). The plasmid was transformed into E. coli BL21 strain.

The bacteria were grown in LB at 310 K overnight. 10 ml of the overnight culture was inoculated into 400 ml fresh LB and grown for 2.5 h at 310 K with shaking, induced with 100 µl 1 M IPTG and incubated for a further 2.5 h. The bacteria were pelleted by centrifugation, resuspended in 20 ml TBS (10 mM Tris–HCl pH 8, 150 mM NaCl) and sonicated using a Heat Systems Inc. sonicator Model XL2020 (5 × 10 s sonication followed by 10 s pause) with a half-inch disruptor horn at output setting 4. The debris was removed by centrifugation. 800 µl of a 50% suspension of glutathione Sepharose 4B beads (Amersham) was added to the supernatant and incubated for 30 min at room temperature. The beads were washed three times with TBS. The pelleted beads were resuspended in 200 µl TBS containing 10 units of thrombin (Sigma) and incubated on a rotating-plate mixer for 14 h at room temperature. The supernatant containing released BipD was collected.

The protein was further purified by gel filtration, injecting about 3 ml at a concentration of 0.5 mg ml−1 onto an S75 16/60 FPLC column (Amersham) running at 1 ml min−1 in 20 mM Tris–HCl pH 6.6 and 20 mM NaCl running buffer. Six 2 ml fractions containing the protein were collected and concentrated by centrifugation in 10 kDa molecular-weight cutoff Vivaspin centrifugal concentrators (VivaScience, Sartorius Group) to a concentration of 10 mg ml−1 at 277 K.

SeMet-labelled BipD was produced by expression in the E. coli methionine-auxotrophic strain B834 (DE3). Cultures were grown in LB media to an A600nm of 0.2, pelleted (15 min, 4000g, 277 K), washed in PBS three times and resuspended in 10 µl SeMet medium without SeMet before being used to inoculate SelenoMet Medium Base containing SelenoMet Nutrient Mix and SeMet stock solution (4 ml per litre; Molecular Dimensions, UK). Cells were grown and induced as described above. SeMet-labelled protein was purified as described above. Full incorporation of selenomethionine was confirmed by mass spectrometry (data not shown).

2.2. Crystallization

Initial crystallization conditions were obtained by sparse-matrix screening (Jancarik & Kim, 1991[Jancarik, J. & Kim, S.-H. (1991). J. Appl. Cryst. 24, 409-411.]) using the sitting-drop vapour-diffusion technique. Drops were prepared by mixing 0.2 µl protein solution (10 mg ml−1, 20 mM Tris pH 6.6, 20 mM NaCl) with 0.2 µl reservoir solution and were equilibrated against 100 µl reservoir solution at 293 K. Initial crystals of BipD grew in two weeks in condition No. 22 of Molecular Dimensions Stura Footprint Screen 1 (1 M trisodium citrate, 10 mM sodium borate pH 8.5). Subsequent optimization yielded diffraction-quality crystals of BipD (Fig. 1[link]) at citrate concentrations in the range 1.0–1.17 M and at pH values between 8.4 and 8.5. Crystals of SeMet-labelled BipD were grown as described above from SeMet-labelled sample. The Pt derivative was obtained by addition of 1 µl reservoir solution saturated with K2PtCl4 to the 0.4 µl crystallization drop; crystals were soaked for 1 d at 293 K prior to data collection.

[Figure 1]
Figure 1
Orthorhombic crystals of native BipD. The largest crystal has dimensions of approximately 20 × 20 × 300 µm.

2.3. Data collection and processing

Crystals of native and SeMet BipD were cryoprotected in reservoir solution containing 20% glycerol and flash-cryocooled in liquid nitrogen for data collection. The crystal of the Pt derivative of BipD was mounted and flash-cryocooled directly in the cryostream.

Diffraction data were recorded at 100 K (Table 1[link]). Data were indexed and integrated in MOSFLM (Leslie, 1992[Leslie, A. G. W. (1992). Jnt CCP4/ESF-EACBM Newsl. Protein Crystallogr. 26.]) and scaled with SCALA (Evans, 1997[Evans, P. R. (1997). Jnt CCP4/ESF-EACBM Newsl. Protein Crystallogr. 33, 22-24.]) within the CCP4 program suite (Collaborative Computational Project, Number 4, 1994[Collaborative Computational Project, Number 4 (1994). Acta Cryst. D50, 760-­763.]). A fluorescence scan measured on the SeMet crystal across the Se edge yielded values of [f'] = −8.0 and [f''] = 6.4 using the program CHOOCH (Evans & Pettifer, 2001[Evans, G. & Pettifer, R. (2001). J. Appl. Cryst. 34, 82-86.]).

Table 1
BipD X-ray diffraction data-collection statistics

Value in parentheses are for the highest resolution shells.

  Native 1 Native 2 SeMet K2PtCl4 soak
X-ray source ESRF ID29 ESRF ID14-2 ESRF ID29 ESRF ID14-2
Detector ADSC scanner ADSC scanner ADSC scanner ADSC scanner
Space group C222 P21212 P21212 P21212
Z 8 8 8 8
Unit-cell parameters        
a (Å) 103.98 136.47 135.90 136.52
b (Å) 122.79 89.84 89.50 89.55
c (Å) 49.17 50.15 49.97 49.29
Wavelength (Å) 0.9778 0.9330 0.9794 0.9330
Resolution limits (Å) 54–2.6 (2.74–2.6) 54–2.7 (2.85–2.7) 43–3.2 (3.37–3.2) 33–2.7 (2.85–2.7)
Completeness (%) 99.8 (99.8) 99.8 (99.8) 100.0 (100.0) 95.2 (95.2)
Anomalous completeness (%) 99.9 (99.9) 95.0 (73.4)
Measured reflections 48047 (6998) 114952 (8942) 67849 (10074) 106839 (9603)
Unique reflections 10006 17569 10590 16408
I/σ(I) 6.1 (1.6) 7.4 (2.4) 3.9 (1.8) 7.7 (2.2)
Multiplicity 4.8 (4.9) 6.5 (3.6) 6.4 (6.7) 6.5 (5.3)
Rmerge 9.4 (42.7) 8.7 (31.8) 16.9 (41.5) 8.0 (33.1)
Ranom 8.3 (3.6) 4.5 (19.8)
Rmerge = 100 × [\textstyle \sum_{h}[\sum_{i}|\langle I(h)\rangle - I(h)_{i}|/][\textstyle \sum_{i}I(h)_{i}]], where I(h)i is the ith observation of reflection h and 〈I(h)〉 is the mean intensity of all observations of h.
Ranom = 100 × [\textstyle \sum_{h}|\langle I^{+}\rangle - \langle I^{-}\rangle |/][\textstyle \sum_{h}(\langle I^{+}\rangle + \langle I^{-}\rangle)], where 〈I+〉 and 〈I〉 are the mean intensities of the Bijvoet pairs for observation h.

3. Results and discussion

Of all the BipD crystals analysed, all of which were grown under the same conditions and all of which had a similar rod shape, only one belonged to space group C222 (labelled Native 1), while all other crystals examined belonged to space group P21212.

The C222 crystal contains one molecule per asymmetric unit, with a Matthews coefficient of 2.4 Å3 Da−1, corresponding to a solvent content of 47% (Matthews, 1968[Matthews, B. W. (1968). J. Mol. Biol. 33, 491-497.]). The P21212 crystals are likely to contain two molecules per asymmetric unit, based on the frequency distribution of Matthews coefficients among the PDB entries at comparable resolution (Kantardjieff & Rupp, 2003[Kantardjieff, K. A. & Rupp, B. (2003). Protein Sci. 12, 1865-1871.]); this is confirmed by the P21212 self-rotation function peak at ω = 90, φ = 27, κ = 180°, which has an intensity equal to 46% of the crystallographic peaks (Fig. 2[link]).

[Figure 2]
Figure 2
The κ = 180° section of the self-rotation function calculated for the P21212 native data set using MOLREP (Collaborative Computational Project, Number 4, 1994[Collaborative Computational Project, Number 4 (1994). Acta Cryst. D50, 760-­763.]) with an integration radius of 27.6 Å and data in the resolution range 37–5 Å. The peak (marked with an X) at (ω, φ) = (90, 27°) represents 46% of the peaks for the crystallographic twofold axes.

The c unit-cell parameter is equal to about 49 Å in both BipD crystal forms; the estimated solvent content and unit-cell volumes are very similar (V = 627 744 and 613 596 Å3 for C222 and P21212, respectively), making it possible that the two lattices are equivalent. Indeed, the cross-crystal rotation function [computed between the C222 and P21212 forms with the program ALMN (Collaborative Computational Project, Number 4, 1994[Collaborative Computational Project, Number 4 (1994). Acta Cryst. D50, 760-­763.]) using data in the resolution range 15–5 Å] has a 5σ peak at ω = 0, φ = 0, κ = 26.7° (data not shown), suggesting that the two lattices are related by a simple rotation around the direction of the c axis. Moreover, this cross-crystal rotation approximately aligns the direction of the a axis of C222 with the NCS twofold in P21212. Thus, the P21212 form is likely to be the result of a change in symmetry that causes one of the twofold axes in C222 to become an NCS twofold in P21212.

Anomalous difference Patterson maps for the Pt and SeMet derivatives were calculated within autoSHARP (Vonrhein et al., 2005[Vonrhein, C., Blanc, E., Roversi, P. & Bricogne, G. (2005). In Crystallographic Methods, edited by S. Doublié. Totowa, NJ, USA: Humana Press.]) using (E2 − 1) coefficients and strict outlier rejection. The Pt anomalous difference Patterson Harker sections suggest the presence of two sites (Fig. 3[link]), while the SeMet anomalous difference Pattersons for the expected 2 × 6 = 12 Se sites are not easily interpretable (data not shown), nor have they yielded a reliable set of Se sites, despite attempts using the available Patterson-solving computer programs. The current phasing strategy is therefore based on initial SIRAS phasing with the Pt sites only, which should enable the location of the subset of ordered Se sites in the SeMet derivative and allow MIRAS phasing with the full BipD native data set and Pt- and SeMet-derivative data sets.

[Figure 3]
Figure 3
Harker section (u = 0.5) of the anomalous difference Patterson maps of the K2PtCl4 soak of BipD (space group P21212) calculated at 3.3 Å resolution with data collected at λ = 0.9330 Å (see Table 1[link]) by autoSHARP. Maps are drawn with a minimum contour level of 1.5σ with 0.3σ increments. The two main peaks are labelled.

Acknowledgements

PR is funded by a Wellcome Trust Grant (No. 077082) to SML and PR. SJ is funded by a grant from the Medical Research Council of the UK (G0400389) to SML. JED is funded by an Australian National Health and Medical Research Council CJ Martin Postdoctoral Fellowship (ID-358785). Work at the IAH is supported by the BBSRC (UK). We are grateful to Ed Lowe and Martin Noble for collecting the diffraction data from the SeMet and K2PtCl4 derivatives. Marc Morgan and Jenny Gibson assisted us with the use of the crystallization robot. Michael Wood was involved in the cloning of BipD.

References

First citationBateman, A., Coin, L., Durbin, R., Finn, R. D., Hollich, V., Griffiths-Jones, S., Khanna, A., Marshall, M., Moxon, S., Sonnhammer, E. L., Studholme, D. J., Yeats, C. & Eddy, S. R. (2004). Nucleic Acids Res. 32, D138–D141.  Web of Science CrossRef PubMed CAS Google Scholar
First citationCollaborative Computational Project, Number 4 (1994). Acta Cryst. D50, 760–­763.  CrossRef IUCr Journals Google Scholar
First citationCordes, F. S., Daniell, S., Kenjale, R., Saurya, S., Picking, W. L., Picking, W. D., Booy, F., Lea, S. M. & Blocker, A. (2005). J. Mol. Biol. 354, 206–211.  Web of Science CrossRef PubMed CAS Google Scholar
First citationDeane, J. E., Cordes, F. S., Roversi, P., Johnson, S., Kenjale, R., Picking, W. D., Picking, W. L., Lea, S. M. & Blocker, A. (2006). Acta Cryst. F62, 302–305.  Web of Science CrossRef IUCr Journals Google Scholar
First citationDeane, J. E., Roversi, P., Cordes, F. S., Johnson, S., Kenjale, R., Daniell, S., Booy, F., Picking, W. D., Picking, W. L., Blocker, A. J. & Lea, S. M. (2006). Proc. Natl Acad. Sci. USA, doi:10.1073/pnas.0602689103.  Google Scholar
First citationDerewenda, U., Mateja, A., Devedjiev, Y., Routzahn, K. M., Evdokimov, A. G., Derewenda, Z. S. & Waugh, D. S. (2004). Structure, 12, 301–306.  Web of Science CrossRef PubMed CAS Google Scholar
First citationEvans, G. & Pettifer, R. (2001). J. Appl. Cryst. 34, 82–86.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEvans, P. R. (1997). Jnt CCP4/ESF–EACBM Newsl. Protein Crystallogr. 33, 22–24.  Google Scholar
First citationJancarik, J. & Kim, S.-H. (1991). J. Appl. Cryst. 24, 409–411.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationJohnson, S., Deane, J. E. & Lea, S. M. (2005). Curr. Opin. Struct. Biol. 15, 700–­707.  Web of Science CrossRef PubMed CAS Google Scholar
First citationKantardjieff, K. A. & Rupp, B. (2003). Protein Sci. 12, 1865–1871.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLeslie, A. G. W. (1992). Jnt CCP4/ESF–EACBM Newsl. Protein Crystallogr. 26Google Scholar
First citationMatthews, B. W. (1968). J. Mol. Biol. 33, 491–497.  CrossRef CAS PubMed Web of Science Google Scholar
First citationMota, L. J. (2006). Trends Microbiol. 14, 197–200.  Web of Science CrossRef PubMed CAS Google Scholar
First citationRotz, L. D., Khan, A. S., Lillibridge, S. R., Ostroff, S. M. & Hughes, J. M. (2002). Emerg. Infect. Dis. 8, 225–230.  CrossRef PubMed Google Scholar
First citationStevens, M. P. & Galyov, E. E. (2004). Int. J. Med. Microbiol. 293, 549–555.  Web of Science CrossRef PubMed CAS Google Scholar
First citationStevens, M. P., Haque, A., Atkins, T., Hill, J., Wood, M. W., Easton, A., Nelson, M., Underwood-Fowler, C., Titball, R. W., Bancroft, G. J. & Galyov, E. E. (2004). Microbiology, 150, 2669–2676.  Web of Science CrossRef PubMed CAS Google Scholar
First citationStevens, M. P., Wood, M. W., Taylor, L. A., Monaghan, P., Hawes, P., Jones, P. W., Wallis, T. S. & Galyov, E. E. (2002). Mol. Microbiol. 46, 649–659.  Web of Science CrossRef PubMed CAS Google Scholar
First citationVonrhein, C., Blanc, E., Roversi, P. & Bricogne, G. (2005). In Crystallographic Methods, edited by S. Doublié. Totowa, NJ, USA: Humana Press.  Google Scholar
First citationWhite, N. J. (2003). Lancet, 361, 1715–1722.  Web of Science CrossRef PubMed CAS Google Scholar
First citationWilkinson, L. (1981). Med. Hist. 25, 363–384.  CrossRef CAS PubMed Web of Science Google Scholar
First citationZhang, L., Wang, Y., Picking, W. L., Picking, W. D. & De Guzman, R. N. (2006). J. Mol. Biol. 359, 322–330.  Web of Science CrossRef PubMed CAS Google Scholar

© International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoSTRUCTURAL BIOLOGY
COMMUNICATIONS
ISSN: 2053-230X
Follow Acta Cryst. F
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds