You are on page 1of 14

Research

Two-dimensional Blue Native/SDS Gel Electrophoresis of Multiprotein Complexes from Helicobacter pylori*
Slove nie Pyndiah, Jean Paul Lasserre, Armelle Me nard, Ste phane Claverol, Vale rie Prouzet-Maule on, Francis Me graud**, Frank Zerbib, and Marc Bonneu
The study of protein interactions constitutes an important domain to understand the physiology and pathogenesis of microorganisms. The two-dimensional blue native/SDSPAGE was initially reported to analyze membrane protein complexes. In this study, both cytoplasmic and membrane complexes of a bacterium, the strain J99 of the gastric pathogen Helicobacter pylori, were analyzed by this method. It was possible to identify 34 different proteins grouped in 13 multiprotein complexes, 11 from the cytoplasm and two from the membrane, either previously reported partially or totally in the literature. Besides complexes involved in H. pylori physiology, this method allowed the description of interactions involving known pathogenic factors such as (i) urease with the heat shock protein GroEL or with the putative ketol-acid reductoisomerase IlvC and (ii) the cag pathogenicity island CagA protein with the DNA gyrase GyrA as well as insight on the partners of TsaA, a peroxide reductase/stress-dependent molecular chaperone. The two-dimensional blue native/ SDS-PAGE combined with mass spectrometry is a potential tool to study the differences in complexes isolated in various situations and also to study the interactions between bacterial and eucaryotic cell proteins. Molecular & Cellular Proteomics 6:193206, 2007. opment of reliable proteome-wide approaches for a better understanding of the virulence mechanisms of the bacterium. One of the strategies of functional proteomics, a method used to identify gene function at the protein level, is the comprehensive analysis of protein-protein interactions related to the functional linkage among proteins and the analysis of functional cellular machinery to better understand the basis of the organisms functions. Recent progress in high throughput technologies has allowed the characterization of protein-protein interactions more directly than ever before using procedures such as the two-hybrid assay (7), co-purification (8, 9), or co-immunoprecipitation (10). The workhorse of experimental proteomics has been the two-hybrid screening method, although it has been criticized for its limited accuracy of results and its laborintensive nature (11, 12). Indeed, it is currently the most reliable technique for large scale characterization of protein interactions in complete genomes (13). Protein chips may eventually provide large scale simultaneous protein-protein interaction data (14, 15), but technical problems (denaturation and substrate biocompatibility) must be overcome to scale up for high throughput analysis. To date, these technologies have generated large interaction networks for bacteria (16), yeast (8, 9), fruit flies (17), and nematode worms (18). Other approaches will undoubtedly become prominent as proteomics technology continues to evolve. The limiting factor for identifying protein complexes is the separation method that must be performed under native conditions to prevent protein dissociation. Because of these limits, the two-dimensional blue native (2D BN)1/SDS-PAGE method was applied to study the H. pylori reference strain J99 complexome. Protein identification was performed by using LC-MS/MS. This highly resolvent separation method was initially described for the separation under native conditions of the membrane protein complexes of mitochondria (19). Later on, numerous

Helicobacter pylori is a spiral, microaerophilic, Gram-negative bacterium that colonizes the gastric epithelium (1) in 40 60% of the worlds population. Approximately 10 20% of these infected individuals suffer from diseases, such as peptic ulcer disease, or from conditions such as chronic atrophic gastritis, which can then evolve toward gastric cancer (2 4). Identification and characterization of multiprotein complexes are important steps in obtaining an integrative view of protein-protein interaction networks that determine protein function and cell behavior. The availability of complete DNA sequences of two H. pylori strains (5, 6) has led to the develFrom INSERM, U853, Bordeaux, F 33076 France and the Laboratoire de Bacte riologie and Po le Prote omique, Plateforme Ge nomique Fonctionnelle Bordeaux, Universite Victor Segalen Bordeaux 2, Bordeaux, F 33076 France Received, September 19, 2006 Published, MCP Papers in Press, November 7, 2006, DOI 10.1074/ mcp.M600363-MCP200

1 The abbreviations used are: 2D BN, two-dimensional blue native; UreA/B, urease / subunit; GroEL, chaperone and heat shock protein; AhpC, alkyl hydroperoxide reductase; TSA, thiol-specific antioxidant; POR, pyruvate oxidoreductase; PAI, pathogenicity island; CagA or -I, cag pathogenicity island protein A or I; GyrA, DNA gyrase subunit A; OMP, outer membrane protein; BabA/B, blood group antigen-binding adhesin A/B.

2007 by The American Society for Biochemistry and Molecular Biology, Inc. This paper is available on line at http://www.mcponline.org

Molecular & Cellular Proteomics 6.2

193

Helicobacter pylori Proteomics

studies focused on the protein complexes of the respiratory chain (20 27). More recently, because this method is reproducible, it was successfully used to study the detection of protein complex deficiencies of mitochondria (28 32). In the same manner, this method was applied to the protein complexes of the chloroplast membranes of cyanobacteria (33) and protein complexes of the mitochondrial and chloroplast membranes of plants (34 41). Moreover a similar method using agarose instead of acrylamide was developed to study protein complexes of approximate molecular mass greater than 1200 kDa (42). The anionic dye used in this method (Coomassie Brilliant Blue G-250) binds to the surface of all proteins, particularly on aromatic residues and on arginines. This binding of a large number of negatively charged dye molecules to proteins facilitates the multiprotein complex migration in a first dimension native electrophoresis (BN-PAGE), and the tendency for protein aggregation is thus reduced considerably. Multiprotein complexes are separated according to their size and shape. Each multiprotein complex is denatured in a second dimension electrophoresis (SDSPAGE), and the protein alignment allows the identification of interactive proteins. Recent modifications have made it possible to apply this method to whole protein complexes (43 45). Indeed this method was recently used to study the Escherichia coli cell envelope complexome (44), human embryonic kidney 293 cells (43), and human platelets (45). A dialysis must be performed on cytoplasmic extracts to eliminate salt and small molecules (43). This method was applied to analyze both H. pylori cytoplasmic and membrane complexes. Purification steps such as liquid IEF or chromatography fractionation and enrichment were used to improve the multiprotein complex separation from the cytoplasm.
EXPERIMENTAL PROCEDURES

Bacterial Growth ConditionsH. pylori strain J99 (ATCC 700824) (6) was used for the experiments. H. pylori cells were cultured for 48 h on Wilkins Chalgren agar (Oxoid Ltd., Hampshire, UK) plates supplemented with 10% human blood and the following antibiotics: 10 mg/ml vancomycin (Lilly France S.A., Fergesheim, France), 2 mg/ml cefsulodin (Takeda France S.A., Puteaux, France), 5 g/ml Fungizone (Bristol-Myers Squibb Co.), and 5 mg/ml trimethoprim (GlaxoSmithKline). A bacterial suspension was grown in brucella broth (BD Biosciences) supplemented with the same antibiotics cited above and with 10% fetal calf serum (Eurobio, Les Ulis, France). The plates were incubated at 37 C under microaerobic conditions (5% O2, 10% CO2, 85% N2), and the broth was incubated in a 1-liter loosely capped container, with a 150 rpm agitation, in a microaerobic atmosphere at 37 C for less than 72 h. Bacteria harvested from two or three agar plates were suspended in 250 ml of brucella broth. The resulting liquid culture showed an approximate bacterial growth of 0.1% (w/v). For the development of the 2D BN/SDS-PAGE applied to H. pylori cytoplasmic extract and to obtain the final results for both the cytoplasmic and the membrane extract, a total of 10 g of H. pylori strain J99 was frozen. Bacterial Lysate, Cytoplasmic, and Membrane PreparationsAll of

the bacteria and sample manipulations were performed at 4 C. Bacteria were harvested from culture by a centrifugation at 2,500 g for 10 min and washed twice in PBS buffer. Bacteria were suspended (v/v) in native extraction buffer A (750 mM 6-amino-n-caproic acid, 50 mM Tris/HCl, pH 7.0, at 4 C) supplemented with a 1 mM final concentration of PMSF and passed three times through a One Shot disruptor at 2 kilobars. The lysate was centrifuged at 6,000 g for 20 min, the supernatant was kept, a 0.2 mg/ml final concentration of DNase I was added, and digestion was carried out for 1 h at 25 C. Then the supernatant was centrifuged at 100,000 g for 30 min at 4 C; the resulting supernatant and pellet contained the cytoplasmic protein and the membrane protein, respectively. For the cytoplasmic sample preparation, the supernatant was filtered with a Miracloth membrane (Calbiochem). and this sample was named fraction I. The concentration of 125150 mg/ml protein was obtained, and the sample appeared limpid. For the membrane sample preparation, a bacterial lysate was centrifuged at 6,000 g for 20 min, and the pellet was resuspended in buffer A and passed three times through a One Shot disruptor at 2 kilobars. The resulting lysate was centrifuged at 100,000 g for 30 min, and the pellet was resuspended in buffer A and centrifuged once more. The washing step was performed three times. The protein extraction was then carried out by resuspending the membrane in 12 ml of buffer A supplemented with 2% -D-dodecyl-n-maltoside detergent (Sigma). This sample was then centrifuged at 100,000 g for 30 min, and the membrane multiprotein complexes contained in the supernatant were separated by 2D BN/SDS-PAGE. Concerning the membrane extract, the detergent used interfered with the Bradford protein dosage. For this reason and to obtain sufficiently resolvent gels, a preliminary experiment was carried out with various sample dilutions directly loaded onto the first dimension of the 2D BN/SDSPAGE gel to determine the quantity of material needed for the electrophoresis. Purification Steps on the Cytoplasmic SampleAll of the steps were carried out at 4 C. Purification steps using physical or chemical properties of the multiprotein complexes constituted an important aspect to confirm protein-protein interactions. Therefore, liquid IEF or exclusion filtration methods were used as purification steps before applying the 2D BN/SDS-PAGE. Liquid IEF purification was used to separate the multiprotein complexes according to their pI in a pH range from 3.5 to 10. An aliquot of fraction I (containing 60 mg of protein) was analyzed in a Rotofor system (Rotofor Prep IEF Cell, Bio-Rad). The protein mixture was prepared according to the manufacturers recommendations before filling the Rotofor chamber. The IEF method produced many protein precipitates in the most abundant protein fractions with a pI of 5 6. Very low protein concentrations were found in the basic fractions although a concentration step with the Centricon kit (Amicon, Inc., Beverly, MA) was used. After the IEF of the protein sample, the resulting fractions were desalted in buffer A using a 5-ml HiTrapTM desalting column (Amersham Biosciences). Multiprotein complexes were recovered in 300-l fractions. Indeed for H. pylori cytoplasmic extracts, a preliminary dialysis is necessary to obtain highly resolvent gels. Here, dialysis was replaced by a desalting step, which allows the elimination of small molecules and salts, as was described for the purification of the human embryonic kidney cell line HEK293 (43). Gel filtration purification was also tested. An aliquot of fraction I (300 l containing 60 mg of protein) was loaded on a SuperdexTM 200 column (Amersham Biosciences). Buffer A was run at a flow rate of 0.3 ml/min using the FPLC Akta (Amersham Biosciences). Multiprotein complexes were recovered in 250-l fractions. The cytoplasmic sample was separated into five peaks of major interest: 1000, 580, 340, 100, and 93 kDa. This method allowed the adaptability of

194

Molecular & Cellular Proteomics 6.2

Helicobacter pylori Proteomics

the 2D BN/SDS-PAGE acrylamide gradient according to the mass of interest of the complexes. The Centricon kit (Amicon, Inc.) with a cutoff of 50 kDa was also used to concentrate the purified samples when the sample concentration was lower than 50 mg/ml. First Dimension: BN-PAGESample preparation and BN-PAGE were carried out as described by Schagger and von Jagow (19) with the following minor modifications. The gel dimension was 22 cm 16.5 cm 1 mm. Separating gels with linear 4 9% acrylamide gradient gels were used. Anode and cathode buffers contained 50 mM Tris, 75 mM glycine, and only the cathode buffer was supplemented with 0.002% Serva blue G (Serva, Heidelberg, Germany). Before loading the sample, 1 l of sample buffer (500 mM 6-amino-n-caproic acid, 5% Serva blue G) was added. The gel was run overnight at 4 C at 1 watt. Thyroglobulin (669 kDa) and bovine serum albumin (66 kDa) were used for each BN-PAGE analysis as molecular weight size standards (Sigma). Different acrylamide gradients were tested for the BN-PAGE to improve the multiprotein complex separation. We found that the best multiprotein complex separation was obtained with a linear 4 9% acrylamide gradient. A certain balance needs to be found to optimize both focalization and separation of complexes with a mass greater than 60 kDa. The 4 9% acrylamide gradient was performed three times on the same sample and two more times on a sample originating from another extraction; this experiment showed that the migration distance of the multiprotein complexes was the same (data not shown). This is proof that the multiprotein complexes of H. pylori maintain their global conformation during the electrophoresis. Therefore, a molecular mass could be attributed to the cytoplasmic complexes based on the two different marker proteins mentioned above. Second Dimension: SDS-PAGEIndividual lanes from BN-PAGE were equilibrated for 5 min in an equilibrating buffer containing 1% SDS (w/v), 0.125 mM Tris/HCl, pH 6.8, and then dipped into equilibrating buffer supplemented with 100 mM dithiothreitol (Acros Organics, Morris Plains, NJ) for 15 min. Individual lanes were subsequently soaked in equilibrating buffer supplemented with 55 mM iodoacetamide (Sigma) for 15 min. An ultimate washing step lasting 5 min was performed in the equilibrating buffer without supplement. Individual lanes were placed on a glass plate at the usual position for stacking gels. After positioning the spacers and covering with the second glass plate, the gel was brought into a vertical position. Then the 10% separating gel mixture was poured. After polymerization the stacking gel mixture was poured. Gel StainingSilver staining was performed using a silver staining kit (Sigma) according to the manufacturers instructions. In-gel Protein DigestionSilver-stained proteins separated by SDS-PAGE were excised and destained using the PROTSIL2 silver staining kit (Sigma) according to the manufacturers instructions. Spots were subsequently washed in distilled water/methanol/acetic acid (47.5:47.5:5) until complete destaining. The solvent mixture was removed and replaced by acetonitrile. After shrinking of the gel pieces, acetonitrile was removed, and gel pieces were dried in a vacuum centrifuge. Gel pieces were rehydrated in 10 ng/l trypsin (Sigma) and 50 mM bicarbonate and incubated overnight at 37 C. The supernatant was removed and stored at 20 C, and the gel pieces were incubated for 15 min in 50 mM bicarbonate at room temperature under rotary shaking. This second supernatant was pooled with the previous one, and a distilled water/acetonitrile/acetic acid (47.5:47.5:5) solution was added to the gel pieces for 15 min. This step was repeated again twice. Supernatants were pooled and concentrated in a vacuum centrifuge to a final volume of 30 l. Digests were finally acidified by the addition of 1.8 l of acetic acid and stored at 20 C. On-line Capillary HPLC Nanospray Ion Trap MS/MS Analysis Peptide mixtures were analyzed by on-line capillary HPLC (LC Pack-

ings, Amsterdam, The Netherlands) coupled to a nanospray LCQ ion trap mass spectrometer (ThermoFinnigan, San Jose, CA). Peptides were separated on a 75-m-inner diameter 15-cm C18 PepMapTM column (LC Packings). The flow rate was set at 200 nl/min. Peptides were eluted using a 550% linear gradient of solvent B for 30 min (solvent A was 0.1% formic acid in 5% acetonitrile, and solvent B was 0.1% formic acid in 80% acetonitrile). The mass spectrometer was operated in positive ion mode at a 2-kV needle voltage and a 38-V capillary voltage. Data acquisition was performed in a data-dependent mode consisting of, alternatively in a single run, a full scan MS over the range m/z 300 2000 and a full scan MS/MS in an exclusion dynamic mode. MS/MS data were acquired using a 3 m/z unit ion isolation window, a 35% relative collision energy, and a 5-min dynamic exclusion duration. Data AnalysisData were analyzed by SEQUEST (ThermoFinnigan, Torrance, CA) against a subset of the National Center for Biotechnology Information (NCBI) database consisting of H. pylori strain J99 and 26695 protein sequences. Carbamidomethylation of cysteines (57) and oxidation of methionines (16) were considered as differential modifications. Only peptides with an Xcorr values greater than 1.5 (single charge), 2 (double charge), and 2.5 (triple charge) were retained. In all cases, Cn values have to be greater than 0.1.
RESULTS

Global Presentation of the ResultsMultiprotein complexes from the cytoplasm were named C, and those from the membrane were named M. Putative multihomooligomeric protein complexes were named H. Certain identifications (noted S1S5) could not be attributed to complexes; they are not presented in Table I but are noted in the legends of the figures only. Proteins that are components of the same multiprotein complex co-migrate in the first dimension and are found aligned with a similar shape in the second dimension. This is the basic condition for the identification of a multiprotein complex. As an example, several spots were found to be perfectly aligned (Figs. 1A and 2), and the five proteins with a similar shape (spots 2125) were attributed to the multiprotein complex named C8. However, the elongated form of these spots in Fig. 1A has a more rounded form in Fig. 2. In a further example, spots 4 6 (Fig. 3) fulfilled these criteria, and the three corresponding proteins were considered to be subunits of a multiprotein complex named C2. Aligned spots but with different shapes were considered to belong to different multiprotein complexes. Obvious examples are indicated by arrows (spots A1A6) in Figs. 2 and 3. The spots A1 and A2 indicated by arrows (Fig. 2) are perfectly aligned; however, spot A1 has a rounded form, whereas spot A2 has a more elongated shape. For this reason, they were attributed to different complexes. Other examples of aligned spots with different shapes are indicated by arrows: 1) spots A3 and A4 in Fig. 2 and 2) spots A5 and A6 in Fig. 3. One can also notice that only spot A6 was identified in the crude extract (Fig. 1A). All these spots (spots A1A6) were considered to belong to distinct complexes. The crude cytoplasmic and membrane protein complexes separated by 2D BN/SDS-PAGE are shown in Fig. 1, A and B, respectively. To increase the number of multiprotein com-

Molecular & Cellular Proteomics 6.2

195

196
Protein component identification Total/partial complexes described previously (Ref.)b n
% kDa

TABLE I Description of the protein complexes identified in H. pylori reference strain J99 using 2D BN/SDS-PAGE Multiprotein complexes named C and M correspond to complexes isolated from the cytoplasm and from the membrane compartments, respectively. Sample preparation before applying the 2D BN/SDS-PAGE is indicated for each protein complex. Putative multihomooligomeric protein complexes were named H. The multiprotein complexes were all localized on 2D BN/SDS-PAGE gels performed in this study and are represented in Figs. 1, 2, and 3. Molecular mass represents the experimental approximate molecular mass. n represents the number of peptides. Cov. represents the protein sequence coverage (percentage) of the peptides. HP, H. pylori; OO, other organisms.

Complex description

Protein complex no. Locus jhp no. Protein annotationa Cov. Xcorr (charge)

Molecular mass

Spot no.

GenBankTM accession no. Theoretical molecular mass

HP

OO

Helicobacter pylori Proteomics

kDa

Molecular & Cellular Proteomics 6.2


jhp0632 jhp0633 jhp0631 jhp1037 jhp1038 jhp1035 gi15612100 2 gi15612103 6 3 8 2.11 (2) 1.53 (1) gi15611699 gi15611700 gi15611698 gi15612102 6 10 3 2 10 15 33 6 86.415 78.148 19.527 3.45 (3) 2.67 (2) 44.599 34.832 20.892 61, 62, 80 jhp1030 jhp0387 jhp1471 jhp1471 jhp0979 jhp0632 jhp0495 jhp0641 jhp0768 gi15611708 gi15611835 gi15611699 gi15611562 gi15612095 gi15611455 gi15612536 gi15612536 gi15612044 4 4 5 5 9 6 8 8 9 18 13 22 22 13 10 12 11 24 38.782 40.694 22.112 22.112 84.968 86.415 129.511 92.457 51.522 jhp0371 jhp0022 jhp0067 jhp0008 gi15611079 gi15611439 gi15611093 gi15611138 Predicted N-methylhydantoinase Predicted N-methylhydantoinase Predicted coding region JHP0631 Pyruvate ferredoxin oxidoreductase, subunit Pyruvate ferredoxin oxidoreductase, subunit Pyruvate ferredoxin oxidoreductase, subunit Predicted mannitol dehydrogenase Predicted proline peptidase Predicted alkyl hydroperoxide reductase Predicted alkyl hydroperoxide reductase Predicted phenylalanyl-tRNA synthetase subunit Predicted N-methylhydantoinase cag pathogenicity island protein A, immunodominant antigen Predicted DNA gyrase subunit A Predicted inosine-5-monophosphate dehydrogenase Predicted coding region JHP0371 Predicted citrate synthase Urease B subunit/urea amidohydrolase 4 8 2 8 2 24 5 17 50.141 48.099 61.509 58.068 16, 49, 50, 81, 82 jhp0068 jhp0067 jhp0008 jhp0313 jhp0068 jhp0991 gi15611139 gi15611138 gi15611079 gi15611381 gi15611139 gi15612056 Chaperone and heat shock protein (protein CPN60, GroEL protein, Hsp60, HspB) Urease A subunit/urea amidohydrolase Urease B subunit/urea amidohydrolase 8 7 10 4 9 2 32 14 21 14 36 7 26.418 61.509 58.068 36.361 26.418 18.119 4.34 (2) 4.33 (2) 16, 49, 50, 81, 82 Chaperone and heat shock protein (protein CPN60, GroEL protein, Hsp60, HspB) Predicted ketol-acid reductoisomerase Urease A subunit/urea amidohydrolase Thiol peroxidase

Multiheterooligomeric complexes Cytoplasm C1c,d 670 1 JHP0632 2 hyuA 3 JPH0631 d,e C2 250 4 porA

porB

porG

C3e

500

C4d

550

7 8 9 10 11

JHP1030 pepQ tsaA tsaA pheT

C5d,e

475

12 13

JHP0632 cagA, cag26

C6d

420

14 15

gyrA guaB

C7

1300

16 17 18

JHP0371 gltA ureB

19

groEl, hspb, hsp60, mopA

C8c,e

1100

20 21

ureA ureB

22

groEL, hspB, hsp60, mopA

23 24 25

ilvC ureA tpx

TABLE I continued jhp1121 jhp0067 jhp0008 jhp0411 jhp0171 3.96 (2) 2.81 (2) 49.256 26.418 25.538 18.119 37.557 60.917 34.832 32.071 2.90 (2) 1.74 (2) 3 6 3 1 15 6 22.068 25.177 97.377 61.509 58.068 16, 49, 50, 81, 82 16, 47, 83, 84 jhp1001 jhp0068 jhp1104 jhp0991 jhp0279 jhp1046 jhp1038 3.01 (2) 2.05 (2) jhp0739 jhp0636 jhp0637 JHP1405 gi15612470 jhp0067 jhp0008 gi15611138 gi15611079 gi15611703 gi15611704 Succinyl-CoA transferase subunit B Succinyl-CoA transferase subunit A 7 2 19 8 gi15611806 4 10 gi15612056 gi15611349 gi15612111 gi15612103 3 20 7 2 9 73 17 10 gi15612066 gi15611139 gi15612169 17 12 4 43 48 27 gi15611479 gi15611241 4 2 7 7 49.868 45.448 gi15611138 gi15611079 14 16 28 30 61.509 58.068 gi15612186 6 3 323.40 16, 49, 50, 81, 82

C9

c,e

900

26

rpoB

27 28

ureB groEL, hspB, hsp60, mopA

29 29

JHP0411 glyA

29 30 31

gdhA ureA deoD

C10c

900

32 33 34 35

tpx amiE ggt porB

C11e

100

36

JHP0739

DNA-directed RNA polymerase - subunit Urease B subunit Chaperone and heat shock protein (protein CPN60, GroEL protein, Hsp60, HspB) Predicted zinc protease Predicted serine hydroxymethyltransferase Predicted glutamate dehydrogenase Urease A subunit/urea amidohydrolase Predicted purine-nucleoside phosphorylase Thiol peroxidase Aliphatic amidase -Glutamyltranspeptidase Pyruvate ferredoxin oxidoreductase, subunit Predicted coding region JHP0739

37 37

scoB, yxjE scoA, yxjD

Membrane M1 900

38

frpB_3, frpB3

39 40

ureB groEL, hspB, hsp60, mopA jhp0068 jhp0178 jhp0177 gi15611247 gi15611139 gi15611248

M2

400

41 42

ureA frdA

9 14 5

36 28 5

26.418 79.934 27.473

48, 8588

43 Putative multihomooligomeric complexes, cytoplasm H1c,d,e 775 H1 jhp1471 jhp0598 jhp0992 jhp0386 jhp1290 gi15612057 gi15611454 gi15612355 gi15611665 gi15612536

frdB

Predicted iron-regulated outer membrane protein Urease B subunit/urea amidohydrolase Chaperone and heat shock protein (protein CPN60, GroEL protein, Hsp60, HspB) Urease A subunit/urea amidohydrolase Fumarate reductase, flavoprotein subunit Fumarate reductase, iron-sulfur subunit

tsaA

Predicted alkyl hydroperoxide reductase Nonheme iron-containing ferritin

5 2 8 4 4

36 10 50 38 26 2.87 (2) 2.39 (2)

22.112 19.170 24.384 18.664 18.067

H2

c,d,e

600

H2

pfr

54, 59, 60, 89, 90 16, 9193 94, 95 9699 100

H3c,e H4c,d,e

300 275

H3 H4

sodB, sodF aroQ, aroD

H5c,e

250

H5

fabZ

Superoxide dismutase Predicted 3-dehydroquinate dehydratase Predicted (3R)-hydroxymyristoyl-(acyl carrier protein) dehydratase

Molecular & Cellular Proteomics 6.2

Helicobacter pylori Proteomics

Gene product/function according to Boneca et al. (101) (genolist pasteur.fr/PyloriGene/genome.cgi). The reference of Rain et al. (16) corresponds to the yeast two-hybrid interactions reported by Hybrigenics. c Protein complex identified when a liquid isoelectrofocalization purification (Rotofor system) was performed before applying 2D BN/SDS-PAGE (Fig. 2). d Protein complex identified when a gel filtration purification (Superdex 200 column) was performed before applying 2D BN/SDS-PAGE (Fig. 3). e Protein complex identified when no purification step was performed before applying 2D BN/SDS-PAGE (Fig. 1).

197

Helicobacter pylori Proteomics

FIG. 1. Analysis of the crude cytoplasmic and membrane samples of H. pylori reference strain J99. The first dimension gel (BN-PAGE) was performed with an acrylamide gradient of 4 9% for the crude cytoplasmic sample (A) and 6 13% for the crude membrane sample (B). C and M represent complexes isolated from the cytoplasm and from the membrane compartments, respectively. H represents putative multihomooligomeric complexes. * represents spots where different proteins were identified (see Table I). A second migration of the C3 and M1 complexes is shown in Boxes 1 and 3, respectively. The migration of the C5 complex identified in Fig. 3 was also performed using the crude cytoplasmic extract, but the quantity of protein loaded on the gel was increased (Box 2). A second migration of spots S3 and S4 is shown in Box 4. Arrow A6 refers to Fig. 3. Two proteins were identified in spot S1: AlpB (or HopB; two peptides, coverage 2%) and AlpA (or HopC; three peptides, coverage 2%). Spot S2 corresponds to TsaA (five peptides, coverage 31%). Spot S3 contains FrpB3 (22 peptides, coverage 30%), CagA (four peptides, coverage 2%), and ClpB (four peptides, coverage 2%), and spot S4 contains BabB (seven peptides, coverage 13%), HydB (or HyaB; 11 peptides, coverage 18%), and HopM (or OMP5, six peptides, coverage 7%). Spots D1 D14 correspond to monomers of AtpA, AtpD, GroEL, Tig, RecA, AddB, Tfs, TufA, JHP0971, JHP1356, Adk, PorC, CagL, and JHP1494, respectively.

plexes observed, non-denaturing purification steps were performed on the cytoplasmic extract before applying the 2D BN/SDS-PAGE separation, e.g. liquid IEF (Fig. 2) or exclusion filtration (Fig. 3). All of the multiprotein complexes identified in this study are described in Table I. Attempts to identify proteins were made on a large number of spots, but because some LC-MS/MS identifications failed, certain complexes have not been reported in this study. Different types of complexes were observed. Multiheterooligomeric complexes fulfilling the previously defined criteria were clearly identified, and in certain cases, these complexes were found several times. Complex C1 (spots 13) is an

example; it is found in Figs. 2 and 3. The aligned spots are elongated in Fig. 3 but have a more marked curve in Fig. 2. This complex is no doubt present in Fig. 1A, but only spots 1 and 2 can be seen. Because spots 1 and 2 are more intensely colored than spot 3 (in Figs. 2 and 3) and the spots 1 and 2 are very light in Fig. 1, a plausible explanation is that the latter could not be visualized in Fig. 1. One can also notice that the shape of spot 3 is more difficult to interpret because the gel migration is diffuse in this molecular mass range. Interestingly the genes jhp0631, jhp0632, and jhp0633 that encode these three proteins are located on the same operon (6). Complex C2 (spots 4 6) is clearly visualized in Fig. 3 but to a lesser

198

Molecular & Cellular Proteomics 6.2

Helicobacter pylori Proteomics

FIG. 2. Protein complexes identified from H. pylori reference strain J99 when the cytoplasmic sample was purified using the liquid isoelectrofocalization method (Rotofor system) before applying 2D BN/SDS-PAGE. Analysis of the sample fraction containing protein complexes with a global pI of 5 is shown. The first dimension gel (BNPAGE) was performed with an acrylamide gradient of 4 9%. H represents putative multihomooligomeric complexes. * represents spots where different proteins were identified (see Table I). Spots S5S7 all correspond to TsaA (spot S5: three peptides, coverage 13%; spot S6: three peptides, coverage 16; and spot S7: five peptides, coverage 22%). Two examples of aligned spots with different shapes are indicated by arrows A1 and A2 and by arrows A3 and A4, respectively.

FIG. 3. Protein complexes identified from H. pylori reference strain J99 when the cytoplasmic sample was purified using the gel filtration method (Superdex 200 column) before applying 2D BN/SDS-PAGE. Analysis of the sample fraction containing protein complexes in the range of 580 kDa is shown. The first dimension gel (BN-PAGE) was performed with an acrylamide gradient of 4 9%. An example of aligned spots with different shapes is indicated by arrows A5 and A6.

extent than in Fig. 1A because spot 4 is not perfectly aligned with spots 5 and 6 in this figure. Indeed a mixture of proteins containing DnaN (23 peptides, coverage 83%) and PorA

(six peptides, coverage 18%) was identified in spot 4 in the crude extract. In fact, spot 4 probably corresponds to two different spots with the spot corresponding to DnaK masking

Molecular & Cellular Proteomics 6.2

199

Helicobacter pylori Proteomics

the less intense spot corresponding to PorA. The complex C3 represented by spots 710 was identified several times in the crude extract (Fig. 1, A and Box 1). However, after the exclusion filtration purification step, not all of the spots comprising the complex were found. Only the most intense spots, i.e. spots 9 and 10, were found at a molecular mass of 500 kDa (Fig. 3). They both correspond to alkyl hydroperoxide reductase TsaA (Table I). After the liquid IEF purification step, only three intense spots were found at a molecular mass of 500 kDa, and they were also all identified as TsaA (Fig. 2, spots S5S7). These results are in agreement with the fact that TsaA has been described as an abundant protein in H. pylori (46). The results also strongly suggest that spots 7 and 8 in complex C3, which were already very light in Fig. 1A, were probably not detected after the different purification steps due to their weak intensity or due to a loss of these two subunits during the purification process. In addition, after the liquid IEF purification step, all of the visible spots on the Fig. 2 gel were analyzed by LC-MS/MS, and the proteins corresponding to spots 7 and 8 were not found in the 75 analyzed spots (data not shown). Among the multiheterooligomeric complexes, certain ones were not identified until after the purification step. The exclusion filtration purification of the samples enabled the identification of complexes C4, C5, and C6 (Fig. 3). The analysis of the sample fraction containing protein complexes eluted in the range of 580 kDa from the exclusion filtration allowed the identification of complex C4, comprised of two subunits. Indeed spots 11 and 12 are aligned and have the same elongated form. Similarly spots 13 and 14 with their lengthened form belong to complex C5 and were also found in the crude cytoplasmic extract when the first dimension gel was performed with a preparation containing more proteins (Fig. 1, A, H, and Box 2). Furthermore this complex was also found in other 2D BN/SDS-PAGE gels (data not shown). Lastly, complex C6 contains three subunits represented by the similar and aligned spots 1517 (Fig. 3). Problems were encountered for certain identifications (Figs. 1A and 2). For example, spots 18 20 from complex C7 were found to be aligned with a similar shape as were spots 2125 from complex C8. Spots 18 20 (UreB, GroEL, and UreA) from complex C7 were also found in complexes C8 and C9, but the quantities of the corresponding proteins appeared lower in complexes C7 and C8 compared with complex C9, although this amount was estimated on a silver-stained gel considered as not adequate to provide a quantitative measure. This may suggest that the C7-specific subunits are present in C8 and C9 but are near or below the detection limit of the silver stain. However, these three complexes migrate at different molecular mass, suggesting different states of oligomerization. Consequently three different complexes were attributed. In addition, spots 26 32 that were found to be aligned and to all have the same shape in Fig. 1A were also identified in Fig. 2 but with the additional spots 3335. These spots also seem to be present after purification by exclusion filtration (Fig. 3); unfor-

tunately their identification after this purification step was not sufficient to attribute spots 26 32 and 3335 either to a single complex or to two different complexes. A hydroxyapatite affinity chromatography purification (HA Ultrogel column) was also performed, but the complex corresponding to spots 26 32 did not show an affinity to calcium and was directly eluted as demonstrated by the 2D BN/SDS-PAGE analysis of this fraction (data not shown). In fact, the analyses by 2D BN/SDS-PAGE and LC-MS/MS of the unretained fraction and the fractions eluted with the 1300 mM linear gradient of sodium phosphate buffer did not allow the confirmation of the presence of all of these spots in only one complex or in two different complexes because spots 26 32 were never identified in these different fractions (data not shown). The absence of spots 3335 on the gel in Fig. 1A can be explained in two ways: 1) either the proteins present in spots 26 32 and 3335 are all part of the same complex and spots 3335 that were less intense were not identified in the crude extract or 2) spots 26 32 belong to a different complex than that of spots 3335 but with the same apparent pI (which would explain why they were not found in the crude extract except after the liquid IEF purification). As a conservative measure to avoid describing a false complex, these spots were attributed to two different complexes, i.e. complexes C9 and C10 containing spots 26 32 and 3335, respectively. Another problem relates to the identification of a mixture of proteins in certain spots, and as a result, the same shape criterion was more difficult to apply. Different complexes can indeed co-migrate in the first dimension, and therefore the spots are aligned on the second dimension. Accordingly complex C11 is comprised of spots 36 and 37 with a protein mixture in spot 37 (Fig. 1A). As a result, it is not possible to evaluate whether the protein present in spot 36 interacts with only one or with two proteins present in spot 37. To the contrary, because the proteins present in spot 37 are the subunits A and B of succinyl-CoA transferase (ScoA and ScoB) of H. pylori and are known to interact together (16, 47), this complex was considered to contain JHP0739, with the two partners ScoA and ScoB (Table I). Furthermore spots H1H5 (Figs. 1A and 2) that were very intense were identified several times. The molecular mass observed during the migration in the first dimension of the complexes corresponding to these spots did not correspond to the molecular mass of the proteins identified in the second dimension. No other similarly formed spot was found in alignment with spots H1H5 either before or after different purifications. Several hypotheses could explain these incoherent migrations in the first and second dimensions: these spots could correspond to multihomooligomeric complexes, or they could also belong to multiheterooligomeric complexes that contain weakly expressed proteins that are not visible. Because these spots were identified several times after different purifications, it can be assumed that they are multihomooligomeric complexes. Furthermore these five complexes

200

Molecular & Cellular Proteomics 6.2

Helicobacter pylori Proteomics

named H1H5 all have already been described as multihomooligomers in other organisms (Table I), and most of the proteins involved in H1H4 complexes, e.g. TsaA, Pfr, SodB, and AroQ, have already been reported in multihomooligomeric protein complexes in H. pylori. For each of these five putative multihomooligomeric protein complexes named H1 H5, the putative number of subunits was estimated to be 35, 31, 12, 14, and 14, respectively. Concerning the crude membrane sample (Fig. 1B), the complex M1 was easily identified with the aligned spots 38 41 having a slightly elongated and oval form. The spots 42 and 43 were aligned; however, it was difficult to compare their form in the second dimension clearly because the molecular mass corresponding to spot 43 is low, and the gel migration is diffuse in this molecular mass range. The co-migration in the first dimension seems to be correct because these two spots were observed several times (data not shown). Moreover, these proteins (spots 42 and 43) have been identified previously as the subunits A and B of the fumarate reductase (FrdA and FrdB) of H. pylori (48). This complex was named M2. We had difficulties identifying another complex. Indeed spots S3 and S4 were aligned and presented a similar, very elongated shape, and they were identified several times in the crude extract (Fig. 1, B and Box 4). However, the molecular mass observed during the first migration (250 kDa) was lower than the molecular mass of all the proteins identified in the second dimension because protein mixtures were observed in these two spots: spot S3 contained the predicted iron-regulated outer membrane protein FrpB_3 (or FrpB3, JHP1405; 97 kDa), the cag pathogenicity island protein A CagA (or Cag26, JHP0495; 129 kDa), and the predicted ATPdependent protease binding subunit/heat shock protein ClpB (JHP0249; 96 kDa); spot S4 contained the predicted adhesin BabB (or OMP19, JHP1164; 76 kDa), the large subunit of the quinone-reactive Ni/Fe-hydrogenase HydB (or HyaB, JHP0575; 64 kDa), and the predicted outer membrane protein HopM (or OMP5, JHP0212; 75 kDa). The difference between the molecular mass of the complex observed in the first dimension or deduced from the migration of the proteins in the second dimension is too important for these proteins to belong to the same complex. These incoherent migrations in the first and second dimensions could be explained by the presence of at least two complexes that co-migrate in the first dimension and whose subunits have the same mass. Given these results, no complex including these proteins, considered to be important in the virulence of H. pylori, can be described yet, and complementary experiments must be carried out. This example clearly reflects the various difficulties of interpretation of the results obtained by the 2D BN/SDS-PAGE technique as well as the frequent difficulty in assigning certain proteins to a complex when a mixture of proteins is present. A total of 13 multiprotein complexes containing 34 different proteins (46 in total) were identified from H. pylori strain J99. Among these complexes, 11 were obtained from the cyto-

TABLE II Urease complex C9 isolated from H. pylori strain J99 cytoplasm Solid and dotted brackets represent direct protein-protein interactions and interactions including an intermediary protein (indicated in italics), respectively. Uvra corresponds to HP0705 and JHP0644 in H. pylori reference strains 26695 and J99, respectively. *, Dunn et al. (49), , Evans et al. (50); , Hybrigenics (Rain et al. (16)); , the intermediary protein reported in this interaction is specific to H. pylori reference strain 26695, and no homolog exists in H. pylori reference strain J99.

plasm, and two were from the membrane. Seven multiprotein complexes identified in this study were described previously (Table I) either totally or partially, confirming the interest of this method to study the H. pylori complexome. As an example, several interactions between the different proteins from the cytoplasmic complex C9 (Figs. 1A and 2) were partially described in the literature (Table II). These interactions included the interaction between the two structural subunits, UreA and UreB, of the well known urease enzyme (49). Moreover, previous studies described the urease enzyme interacting with GroEL (Hsp60), a chaperone and heat shock protein (50). On the other hand, some protein interactions described in the complex C9 and reported by Hybrigenics (Rain et al. (16)) involved an intermediary protein (Table II) not identified in the current study, e.g. the interaction between GroEL and RpoB included the predicted outer membrane protein JHP0600. This cytoplasmic complex C9 that includes nine different proteins contained the greatest number of interacting proteins observed in this study. In addition to heterooligomeric complexes, five putative multihomooligomeric protein complexes were identified from the cytoplasm (H1H5, Table II). Concerning the membrane, very few complexes are reported in this study due to a poor resolution and/or visualization of the spots observed on the numerous gels performed, although the 2D BN/SDS-PAGE was initially reported for the separation of membrane protein complexes (51). This suggests that membrane purifications should also improve the gel quality. Different oligomerization states were sometimes observed in certain complexes. For example, TsaA was found in different oligomerization states and in both cytoplasmic and membrane samples: C3 and spots H1 and S2 (Figs. 1, A and B; 2; and 3 and Table I). These results are in agreement with those

Molecular & Cellular Proteomics 6.2

201

Helicobacter pylori Proteomics

of Backert et al. (46) who also identified TsaA, both in structure-bound and in soluble fractions of H. pylori, in eight oligomerization states in their 2D gel electrophoresis and mass spectrometry study. In addition, different migrations of TsaA were observed during the separation in the second dimension (Fig. 2, spots S5 S7; and Fig. 3, spots 9 and 10) suggesting that multiple isoforms of TsaA exist. Their occurrence can be explained by probable post-translational modifications modifying the physicochemical criteria (pI, molecular mass, and binding affinity) of the protein, as reported previously for numerous proteins (52, 53), thus altering the migration in the second dimension. In fact, H. pylori proteins are subject to a high degree of post-translational modification, and TsaA is among several proteins present in multiple isoforms in the bacterium (54). TsaA, originally identified as a 26-kDa antigen, belongs to a family of antioxidants called AhpC/TSA (alkyl hydroperoxide reductase/thiol-specific antioxidant) involved in the defense against oxidative stress and survival of the bacterium in the host. More recently, TsaA was shown to switch from a peroxide reductase to a stress-dependent molecular chaperone function (55). H. pylori TsaA was initially reported to induce an immune response in H. pylori-infected patients with gastric adenocarcinoma (56) and also an immune response in H. pylori-infected Mongolian gerbils that is associated with the emergence of gastric diseases such as chronic active gastritis, gastric ulcer, and gastric cancer (57). TsaA belongs to the five immunodominant H. pylori proteins (58). Pentameric arrangements of homodimers of TsaA were described in H. pylori cytoplasm (59, 60). In our study, TsaA was found in homooligomeric form but also in a complex including proteins JHP1030 and PepQ. The presence of JHP1030 with TsaA is not really surprising because JHP1030 is a predicted zinc-dependent mannitol dehydrogenase that possesses an oxidoreductase activity (inferred from electronic annotation). On the other hand, the presence of PepQ in this complex is more difficult to interpret because it is a predicted proline peptidase that has high homology with prolidases and X-prolyl dipeptidyl aminopeptidases and can be involved either in protein maturation or in nitrogen metabolism. Other experiments should make it possible to determine the exact role of these two partners of TsaA. Complexes Involved in H. pylori VirulenceBefore considering virulence factors stricto sensu, complexes containing proteins involved in the energy metabolism (electron transport pathway, tricarboxylic acid cycle, and gluconeogenesis), which is a prerequisite for virulence, will be briefly presented. Three pyruvate ferredoxin oxidoreductases, PorA, PorB, and PorC (ex-PorG), involved in electron transport were found together in complex C2 (Table I and Figs. 1A and 3). The heterotetrameric POR complex is comprised of the subunits PorA, PorB, PorD, and PorC (61) and was reported previously to be essential because inactivation of porB appeared to be lethal for H. pylori (62). PorD, the fourth subunit of the POR complex, was not found in this study probably because its

molecular mass (14.9 kDa) is near the border limit of the gel migration. POR has been implicated in the bioreduction of nitroimidazole drugs, particularly metronidazole used for H. pylori eradication (61, 63). To understand the mechanisms of metronidazole activation and resistance, which are currently poorly characterized, a more detailed knowledge of the electron transport pathways of the enzymes involved would be helpful. Fumarate reductase catalyzes the reduction of fumarate to succinate in the Krebs cycle and appears to play an important role in the energy metabolism of H. pylori. The two catalytic subunits of H. pylori fumarate reductase, FrdA and FrdB (48), were found in the membrane complex M2 (Fig. 1B and Table I). Fumarate reductase is not necessary for H. pylori survival in vitro but is essential for H. pylori colonization of the mouse stomach (64). It was shown to be strongly immunogenic in sera from H. pylori-positive patients, suggesting that this protein involved in H. pylori energy metabolism could also be used as an anti-H. pylori vaccine candidate (65). Examples of multiprotein complexes including known H. pylori virulence factors are described below. The Urease EnzymeH. pylori produces large amounts of urease that catalyzes the hydrolysis of urea to yield ammonia and carbon dioxide, neutralizing hydrogen ions before they can lower the intracellular pH, thus enabling the bacterium to survive in gastric acid and to colonize the gastric mucosa (66). This urease activity is thus required for colonization in vivo (67). The native urease complex was found in both the membrane and the cytoplasm of the bacteria. Different oligomerization forms of the urease were identified in the cytoplasm (Table I). Four urease complexes named C7, C8, C9, and M1 with a global pI of 5 and a molecular mass range of 900 1300 kDa were identified with crude and purified extracts (Figs. 1, A and B, and 2 and Table I). In fact, UreA and UreB are known to be present in multiple isoforms in H. pylori (54) and are among the predominant H. pylori proteins identified during cell infection (68). H. pylori is known to produce a 550-kDa heterohexameric enzyme composed of three (UreA) and three subunits (UreB). Four heterohexamers (2200 kDa) form a 16-nm spherical complex also produced by the bacteria (69). This kind of oligomerization may be adapted for acid resistance. The separation resolution of the 2D BN/SDSPAGE did not allow the visualization of large complexes with a molecular mass greater than 1500 kDa. All of the urease complexes identified in this study included the UreA/UreB/ GroEL core. These three proteins were also reported to be present in structure-bound and in soluble fractions of H. pylori (46) and were strongly reactive to specific antibodies (58), indicating that this core is highly immunodominant in H. pylori. Under all pH conditions, the most abundant proteins observed were the urease structural subunit UreB and the chaperonin GroEL (70). A close interaction between GroEL, the co-chaperone protein GroES, and the urease enzyme had been suspected (50, 71), but until now no interaction between

202

Molecular & Cellular Proteomics 6.2

Helicobacter pylori Proteomics

GroEL and the urease has been confirmed. The C8, C9, and M1 complexes included additional proteins besides the core UreA/UreB/GroEL. The C8 complex included the IlvC protein, a putative ketol-acid reductoisomerase involved in isoleucinevaline biosynthesis and the thiol peroxidase Tpx involved in detoxification; the M1 complex included the FrpB_3 protein, a putative iron-regulated outer membrane protein; and the C9 complex included six other proteins (Tables I and II). cag Pathogenicity Island (PAI) ProteinsCagA is the marker and one of the effectors of the cagPAI, translocated in gastric epithelial cells and conferring increased virulence to H. pylori strains (72). CagA was identified both in the cytoplasm and in the membrane: complex C5 and spot S3, respectively (Figs. 1, A and B, and 3 and Table I). This localization of CagA in the membrane was reported previously (46) and is probably due to its presence in the type IV secretion system in the H. pylori cell wall (73, 74). The multiprotein complex C5 including CagA and the DNA gyrase subunit A (GyrA) was found in the cytoplasmic sample with a molecular mass of 475 kDa (Figs. 1A and 3) suggesting a probable oligomerization of CagA and/or GyrA in this complex. This CagA-GyrA complex was also observed on other 2D BN/SDS-PAGE (data not shown). Cultures of CagA-negative H. pylori strains show a slower growth than the CagA-positive strains (72, 75). This phenomenon could be explained by the CagA/GyrA interaction described in this study because this interaction may have a favorable effect on the normal DNA replication process, leading to a better development of the bacterial cell. However, Hybrigenics (Rain et al. (16)) described an interaction between GyrA and the cagPAI protein I (CagI), including a small intermediary cysteine-rich protein B (HcpB), which is a weak -lactamase.
DISCUSSION

The classical methods widely used for large scale analysis of protein interactions are the yeast two-hybrid and the tandem affinity protein tag methods with the main limitation of expressing chimerical proteins. With the yeast two-hybrid method, mostly binary interactions are identified. Furthermore the proteins of interest are often expressed in a heterologous system, the yeast, and the interactions are observed in a particular compartment, the nucleus. Moreover the reliability of high throughput yeast two-hybrid assays is 50% (76). In the tandem affinity protein tag method, the protein complexes are analyzed in vivo, but the organism being studied needs to be competent for exogenous DNA. Another disadvantage of this method is that multihomooligomeric complexes cannot be analyzed. The 2D BN/SDS-PAGE used for the study of the H. pylori complexome also has some limitations. First, it can be difficult to easily assign each spot to independent protein complexes when they are located on the same vertical line in the second dimension; complexes C8 and C9 are very representative of this problem as well as spots containing a mixture of proteins

such as spots S3 and S4 for which no complex could be attributed. For this reason, it is sometimes necessary not only to analyze the crude extract but also to carry out various methods of purifications based on different physicochemical criteria (pI-, molecular mass-, or affinity-based purification). This makes it possible to validate the identified complex and to identify complexes sometimes with a better resolution and/or visualization. However, subunits of unstable complexes can be lost during these various purification steps, therefore leading to the description of incomplete complexes. Second, proteins that are weakly expressed in certain complexes may not be visible, again leading to the description of incomplete complexes or to false multihomooligomeric complexes. In addition, mixtures of proteins can sometimes be identified in the same spots, making it difficult to determine whether all of the identified proteins belong to the same complex. This was a frequently found case for the membrane samples analyzed in this study where the outer membrane proteins often had similar molecular mass and pI. For this reason, several complexes were not reported in this study. Third, the extraction is carried out under non-denaturing conditions, and therefore the complexes requiring denaturing conditions for extraction are not easily found. Examples include 1) H. pylori vacuolating cytotoxin VacA, which is quickly exported over the membrane (77, 78) but was not detected with this method, and 2) flagellar sheaths, anchored in the membrane and requiring denaturing extraction conditions (79), which were not detected either. However, the use of non-denaturing conditions is, in fact, the major advantage of this technique because it provides the native form of protein complexes and thus allows an analysis of the physiologic state of the organism. The method was highly reproducible, and even if it is not possible to determine the totality of the H. pylori complexes, this technique is of great interest to compare the complexes of a bacterium in two different physiological states. Moreover immunoblot applied to the 2D BN/SDS-PAGE would lead to the identification of specific protein complexes, particularly those including strain-specific proteins and virulence factors. Therefore, the comparison of the currently available results obtained with all of the different methods used to analyze protein complexes will allow an accurate identification of the H. pylori protein-protein interaction network.
AcknowledgmentWe are grateful to Corinne Asencio for technical support in bacterial cell culture. * This work was supported in part by the Conseil Re gional dAquitaine and the Association pour la Recherche sur le Cancer. The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked advertisement in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. Recipient of a predoctoral fellowship from the Fondation pour la Recherche Me dicale. ** To whom correspondence should be addressed: Universite Victor Segalen Bordeaux 2, Laboratoire de Bacte riologie, Bat. 2B

Molecular & Cellular Proteomics 6.2

203

Helicobacter pylori Proteomics

RDC Zone Nord, 33076 Bordeaux cedex, France. Tel.: 33-5-56-79-5910; Fax: 33-5-56-79-60-18; E-mail: francis.megraud@chu-bordeaux.fr.
REFERENCES
1. Warren, J. R., and Marshall, B. (1983) Unified curved bacilli on gastric epithelium in active chronic gastritis. Lancet 1, 12731275 2. Correa, P., and Chen, V. W. (1994) Gastric cancer. Cancer Surveys 19 20, 5576 3. Blaser, M. J. (1995) The role of Helicobacter pylori in gastritis and its progression to peptic ulcer disease. Alimentary Pharmacology and Therapeutics 9, 2730 4. Sipponen, P., Hyva rinen, H., Seppa la , K., and Blaser, M. J. (1998) Review article: pathogenesis of the transformation from gastritis to malignancy. Aliment. Pharmacol. Ther. 12, 6171 5. Tomb, J. F., White, O., Kerlavage, A. R., Clayton, R. A., Sutton, G. G., Fleischmann, R. D., Ketchum, K. A., Klenk, H. P., Gill, S., Dougherty, B. A., Nelson, K., Quackenbush, J., Zhou, L., Kirkness, E. F., Peterson, S., Loftus, B., Richardson, D., Dodson, R., Khalak, H. G., Glodek, A., McKenney, K., Fitzegerald, L. M., Lee, N., Adams, M. D., Hickey, E. K., Berg, D. E., Gocayne, J. D., Utterback, T. R., Peterson, J. D., Kelley, J. M., Cotton, M. D., Weidman, J. M., Fujii, C., Bowman, C., Watthey, L., Wallin, E., Hayes, W. S., Borodovsky, M., Karp, P. D., Smith, H., O., Fraser, C. M., and Venter, J. C. (1997) The complete genome sequence of the gastric pathogen Helicobacter pylori. Nature 388, 539 547 6. Alm, R. A., Ling, L. S. L., Moir, D. T., King, B. L., Brown, E. D., Doig, P. C., Smith, D. R., Noonan, B., Guild, B. C., deJonge, B. L., Carmel, G., Tummino, P. J., Caruso, A., UriaNickelsen, M., Mills, D. M., Ives, C., Gibson, R., Merberg, D., Mills, S. D., Jiang, Q., Taylor, D. E., Vovis, G. F., and Trust, T. J. (1999) Genomic-sequence comparison of two unrelated isolates of the human gastric pathogen Helicobacter pylori. Nature 397, 176 180 7. Fields, S., and Song, O. (1989) A novel genetic system to detect proteinprotein interactions. Nature 340, 245246 8. Uetz, P., and Hughes, R. E. (2000) Systematic and large-scale two-hybrid screens. Curr. Opin. Microbiol. 3, 303308 9. Gavin, A. C., Bosche, M., Krause, R., Grandi, P., Marzioch, M., Bauer, A., Schultz, J., Rick, J. M., Michon, A. M., Cruciat, C. M., Remor, M., Hofert, C., Schelder, M., Brajenovic, M., Ruffner, H., Merino, A., Klein, K., Hudak, M., Dickson, D., Rudi, T., Gnau, V., Bauch, A., Bastuck, S., Huhse, B., Leutwein, C., Heurtier, M. A., Copley, R. R., Edelmann, A., Querfurth, E., Rybin, V., Drewes, G., Raida, M., Bouwmeester, T., Bork, P., Seraphin, B., Kuster, B., Neubauer, G., and Superti-Furga, G. (2002) Functional organization of the yeast proteome by systematic analysis of protein complexes. Nature 415, 141147 10. Aebersold, R., and Mann, M. (2003) Mass spectrometry-based proteomics. Nature 422, 198 207 11. Enright, A. J., Iliopoulos, I., Kyrpides, N. C., and Ouzounis, C. A. (1999) Protein interaction maps for complete genomes based on gene fusion events. Nature 402, 86 90 12. Ito, T., Tashiro, K., and Kuhara, T. (2001) Systematic analysis of Saccharomyces cerevisiae genome: gene network and protein-protein interaction network. Tanpakushitsu Kakusan Koso 46, 24072413 13. Legrain, P., and Selig, L. (2000) Genome-wide protein interaction maps using two-hybrid systems. FEBS Lett. 480, 3236 14. MacBeath, G., and Schreiber, S. L. (2000) Printing proteins as microarrays for high-throughput function determination. Science 289, 1760 1763 15. Hynes, S. O., McGuire, J., and Wadstrom, T. (2003) Potential for proteomic profiling of Helicobacter pylori and other Helicobacter spp. using a ProteinChip array. FEMS Immunol. Med. Microbiol. 36, 151158 16. Rain, J. C., Selig, L., De Reuse, H., Battaglia, V., Reverdy, C., Simon, S., Lenzen, G., Petel, F., Wojcik, J., Scha chter, V., Chemana, Y., Labigne, A., and Legrain, P. (2001) The protein-protein interaction map of Helicobacter pylori. Nature 409, 211215 17. Giot, L., Bader, J. S., Brouwer, C., Chaudhuri, A., Kuang, B., Li, Y., Hao, Y. L., Ooi, C. E., Godwin, B., Vitols, E., Vijayadamodar, G., Pochart, P., Machineni, H., Welsh, M., Kong, Y., Zerhusen, B., Malcolm, R., Varrone, Z., Collis, A., Minto, M., Burgess, S., McDaniel, L., Stimpson, E., Spriggs, F., Williams, J., Neurath, K., Ioime, N., Agee, M., Voss, E., Furtak, K., Renzulli, R., Aanensen, N., Carrolla, S., Bickelhaupt, E., Lazovatsky, Y., DaSilva, A., Zhong, J., Stanyon, C. A., Finley, R. L., Jr., White, K. P., Braverman, M., Jarvie, T., Gold, S., Leach, M., Knight, J., 18.

19.

20.

21.

22. 23.

24.

25.

26. 27.

28.

29.

30.

31.

32.

Shimkets, R. A., McKenna, M. P., Chant, J., and Rothberg, J. M. (2003) A protein interaction map of Drosophila melanogaster. Science 302, 17271736 Li, S., Armstrong, C. M., Bertin, N., Ge, H., Milstein, S., Boxem, M., Vidalain, P. O., Han, J. D., Chesneau, A., Hao, T., Goldberg, D. S., Li, N., Martinez, M., Rual, J. F., Lamesch, P., Xu, L., Tewari, M., Wong, S. L., Zhang, L. V., Berriz, G. F., Jacotot, L., Vaglio, P., Reboul, J., HirozaneKishikawa, T., Li, Q., Gabel, H. W., Elewa, A., Baumgartner, B., Rose, D. J., Yu, H., Bosak, S., Sequerra, R., Fraser, A., Mango, S. E., Saxton, W. M., Strome, S., Van Den Heuvel, S., Piano, F., Vandenhaute, J., Sardet, C., Gerstein, M., Doucette-Stamm, L., Gunsalus, K. C., Harper, J. W., Cusick, M. E., Roth, F. P., Hill, D. E., and Vidal, M. (2004) A map of the interactome network of the metazoan C. elegans. Science 303, 540 543 Schagger, H., and von Jagow, G. (1991) Blue native electrophoresis for isolation of membrane protein complexes in enzymatically active form. Anal. Biochem. 199, 223231 Reinheckel, T., Wiswedel, I., Noack, H., Augustin, W., Genova, M. L., Bianchi, C., Lenaz, G., Brookes, P. S., Pinner, A., Ramachandran, A., Coward, L., Barnes, S., Kim, H., Darley-Usmar, V. M., Nijtmans, L. G., Henderson, N. S., Holt, I. J., Schagger, H., Pfeiffer, K., Lindsay, J. G., Lamantea, E., Zeviani, M., Singh, P., Jansch, L., Braun, H. P., Schmitz, U. K., Coenen, M. J., van den Heuvel, L. P., Morava, E., Marquardt, I., Girschick, H. J., Trijbels, F. J., Grivell, L. A., Smeitink, J. A., Kruft, V., Klement, P., Van den Bogert, C., and Houstek, J. (1995) Electrophoretic evidence for the impairment of complexes of the respiratory chain during iron/ascorbate induced peroxidation in isolated rat liver mitochondria. Biochim. Biophys. Acta 1239, 4550 Klement, P., Nijtmans, L. G., Van den Bogert, C., and Houstek, J. (1995) Analysis of oxidative phosphorylation complexes in cultured human fibroblasts and amniocytes by blue-native-electrophoresis using mitoplasts isolated with the help of digitonin. Anal. Biochem. 231, 218 224 Schagger, H. (2001) Blue-native gels to isolate protein complexes from mitochondria. Methods Cell Biol. 65, 231244 Schagger, H., and Pfeiffer, K. (2001) The ratio of oxidative phosphorylation complexes IV in bovine heart mitochondria and the composition of respiratory chain supercomplexes. J. Biol. Chem. 276, 3786137867 Nijtmans, L. G., Henderson, N. S., and Holt, I. J. (2002) Blue native electrophoresis to study mitochondrial and other protein complexes. Methods 26, 327334 Brookes, P. S., Pinner, A., Ramachandran, A., Coward, L., Barnes, S., Kim, H., and Darley-Usmar, V. M. (2002) High throughput two-dimensional blue-native electrophoresis: a tool for functional proteomics of mitochondria and signaling complexes. Proteomics 2, 969 977 Genova, M. L., Bianchi, C., and Lenaz, G. (2003) Structural organization of the mitochondrial respiratory chain. Ital. J. Biochem. 52, 58 61 Navet, R., Jarmuszkiewicz, W., Douette, P., Sluse-Goffart, C. M., and Sluse, F. E. (2004) Mitochondrial respiratory chain complex patterns from Acanthamoeba castellanii and Lycopersicon esculentum: comparative analysis by BN-PAGE and evidence of protein-protein interaction between alternative oxidase and complex III. J. Bioenerg. Biomembr. 36, 471 479 Van Coster, R., Smet, J., George, E., De Meirleir, L., Seneca, S., Van Hove, J., Sebire, G., Verhelst, H., De Bleecker, J., Van Vlem, B., Verloo, P., and Leroy, J. (2001) Blue native polyacrylamide gel electrophoresis: a powerful tool in diagnosis of oxidative phosphorylation defects. Pediatr. Res. 50, 658 665 Coenen, M. J., van den Heuvel, L. P., Nijtmans, L. G., Morava, E., Marquardt, I., Girschick, H. J., Trijbels, F. J., Grivell, L. A., and Smeitink, J. A. (1999) SURFEIT-1 gene analysis and two-dimensional blue native gel electrophoresis in cytochrome c oxidase deficiency. Biochem. Biophys. Res. Commun. 265, 339 344 Wen, J. J., and Garg, N. (2004) Oxidative modification of mitochondrial respiratory complexes in response to the stress of Trypanosoma cruzi infection. Free Radic. Biol. Med. 37, 20722081 Pineau, B., Mathieu, C., Gerard-Hirne, C., De Paepe, R., and Chetrit, P. (2005) Targeting the NAD7 subunit to mitochondria restores a functional complex I and a wild type phenotype in the Nicotiana sylvestris CMS II mutant lacking nad7. J. Biol. Chem. 280, 25994 26001 Perales, M., Eubel, H., Heinemeyer, J., Colaneri, A., Zabaleta, E., and Braun, H. P. (2005) Disruption of a nuclear gene encoding a mitochon-

204

Molecular & Cellular Proteomics 6.2

Helicobacter pylori Proteomics

33.

34.

35.

36.

37.

38.

39.

40.

41. 42.

43.

44.

45.

46.

47.

48.

49.

50.

51.

52.

drial gamma carbonic anhydrase reduces complex I and supercomplex I III2 levels and alters mitochondrial physiology in Arabidopsis. J. Mol. Biol. 350, 263277 Neff, D., and Dencher, N. A. (1999) Purification of multisubunit membrane protein complexes: isolation of chloroplast FoF1-ATP synthase, CFo and CF1 by blue native electrophoresis. Biochem. Biophys. Res. Commun. 259, 569 575 Jansch, L., Kruft, V., Schmitz, U. K., and Braun, H. P. (1996) New insights into the composition, molecular mass and stoichiometry of the protein complexes of plant mitochondria. Plant J. 9, 357368 Singh, P., Jansch, L., Braun, H. P., and Schmitz, U. K. (2000) Resolution of mitochondrial and chloroplast membrane protein complexes from green leaves of potato on blue-native polyacrylamide gels. Indian J. Biochem. Biophys. 37, 59 66 Eubel, H., Jansch, L., and Braun, H. P. (2003) New insights into the respiratory chain of plant mitochondria. Supercomplexes and a unique composition of complex II. Plant Physiol. 133, 274 286 Eubel, H., Heinemeyer, J., and Braun, H. P. (2004) Identification and characterization of respirasomes in potato mitochondria. Plant Physiol. 134, 1450 1459 Eubel, H., Heinemeyer, J., Sunderhaus, S., and Braun, H. P. (2004) Respiratory chain supercomplexes in plant mitochondria. Plant Physiol. Biochem. 42, 937942 Herranen, M., Battchikova, N., Zhang, P., Graf, A., Sirpio, S., Paakkarinen, V., and Aro, E. M. (2004) Towards functional proteomics of membrane protein complexes in Synechocystis sp. PCC 6803. Plant Physiol. 134, 470 481 Millar, A. H., Eubel, H., Jansch, L., Kruft, V., Heazlewood, J. L., and Braun, H. P. (2004) Mitochondrial cytochrome c oxidase and succinate dehydrogenase complexes contain plant specific subunits. Plant Mol. Biol. 56, 7790 Eubel, H., Braun, H. P., and Millar, A. H. (2005) Blue-native PAGE in plants: a tool in analysis of protein-protein interactions. Plant Methods 1, 113 Henderson, N. S., Nijtmans, L. G., Lindsay, J. G., Lamantea, E., Zeviani, M., and Holt, I. J. (2000) Separation of intact pyruvate dehydrogenase complex using blue native agarose gel electrophoresis. Electrophoresis 21, 29252931 Camacho-Carvajal, M. M., Wollscheid, B., Aebersold, R., Steimle, V., and Schamel, W. W. (2004) Two-dimensional blue native/SDS gel electrophoresis of multi-protein complexes from whole cellular lysates: a proteomics approach. Mol. Cell. Proteomics 3, 176 182 Stenberg, F., Chovanec, P., Maslen, S. L., Robinson, C. V., Ilag, L. L., von Heijne, G., and Daley, D. O. (2005) Protein complexes of the Escherichia coli cell envelope. J. Biol. Chem. 280, 34409 34419 Claeys, D., Geering, K., and Meyer, B. J. (2005) Two-dimensional blue native/sodium dodecyl sulfate gel electrophoresis for analysis of multimeric proteins in platelets. Electrophoresis 26, 1189 1199 Backert, S., Kwok, T., Schmid, M., Selbach, M., Moese, S., Peek, R. M., Jr., Konig, W., Meyer, T. F., and Jungblut, P. R. (2005) Subproteomes of soluble and structure-bound Helicobacter pylori proteins analyzed by two-dimensional gel electrophoresis and mass spectrometry. Proteomics 5, 13311345 Corthesy-Theulaz, I. E., Bergonzelli, G. E., Henry, H., Bachmann, D., Schorderet, D. F., Blum, A. L., and Ornston, L. N. (1997) Cloning and characterization of Helicobacter pylori succinyl CoA:acetoacetate CoAtransferase, a novel prokaryotic member of the CoA-transferase family. J. Biol. Chem. 272, 25659 25667 Birkholz, S., Knipp, U., Lemma, E., Kroger, A., and Opferkuch, W. (1994) Fumarate reductase of Helicobacter pylorian immunogenic protein. J. Med. Microbiol. 41, 56 62 Dunn, B. E., Campbell, G. P., Perez-Perez, G. I., and Blaser, M. J. (1990) Purification and characterization of urease from Helicobacter pylori. J. Biol. Chem. 265, 9464 9469 Evans, D. J., Jr., Evans, D. G., Engstrand, L., and Graham, D. Y. (1992) Urease-associated heat shock protein of Helicobacter pylori. Infect. Immun. 60, 21252127 Schagger, H., and von Jagow, G. (1987) Tricine-sodium dodecyl sulfatepolyacrylamide gel electrophoresis for the separation of proteins in the range from 1 to 100 kDa. Anal. Biochem. 166, 368 379 LeHoux, J. G., Fleury, A., Ducharme, L., and Hales, D. B. (2004) Phosphorylation of the hamster adrenal steroidogenic acute regulatory pro-

53.

54.

55.

56.

57.

58.

59.

60.

61.

62.

63.

64.

65. 66.

67.

68.

69.

70.

71.

tein as analyzed by two-dimensional polyacrylamide gel electrophoreses. Mol. Cell. Endocrinol. 215, 127134 Zhang, W., Bergamaschi, D., Jin, B., and Lu, X. (2005) Posttranslational modifications of p27kip1 determine its binding specificity to different cyclins and cyclin-dependent kinases in vivo. Blood 105, 36913698 Lock, R. A., Cordwell, S. J., Coombs, G. W., Walsh, B. J., and Forbes, G. M. (2001) Proteome analysis of Helicobacter pylori: Major proteins of type strain NCTC 11637. Pathology 33, 365374 Chuang, M. H., Wu, M. S., Lo, W. L., Lin, J. T., Wong, C. H., and Chiou, S. H. (2006) The antioxidant protein alkylhydroperoxide reductase of Helicobacter pylori switches from a peroxide reductase to a molecular chaperone function. Proc. Natl. Acad. Sci. U. S. A. 103, 25522557 Wang, J. T., Chang, C. S., Lee, C. Z., Yang, J. C., Lin, J. T., and Wang, T. H. (1998) Antibody to a Helicobacter pylori species specific antigen in patients with adenocarcinoma of the stomach. Biochem. Biophys. Res. Commun. 244, 360 363 Yan, J., Kumagai, T., Ohnishi, M., Ueno, I., and Ota, H. (2001) Immune response to a 26-kDa protein, alkyl hydroperoxide reductase, in Helicobacter pylori-infected Mongolian gerbil model. Helicobacter 6, 274 282 Shin, J. H., Nam, S. W., Kim, J. T., Yoon, J. B., Bang, W. G., and Roe, I. H. (2003) Identification of immunodominant Helicobacter pylori proteins with reactivity to H-pylori-specific egg-yolk immunoglobulin. J. Med. Microbiol. 52, 217222 Krah, A., Schmidt, F., Becher, D., Schmid, M., Albrecht, D., Rack, A., Buttner, K., and Jungblut, P. R. (2003) Analysis of automatically generated peptide mass fingerprints of cellular proteins and antigens from Helicobacter pylori 26695 separated by two-dimensional electrophoresis. Mol. Cell. Proteomics 2, 12711283 Papinutto, E., Windle, H. J., Cendron, L., Battistutta, R., Kelleher, D., and Zanotti, G. (2005) Crystal structure of alkyl hydroperoxide-reductase (AhpC) from Helicobacter pylori. Biochim. Biophys. Acta 1753, 240 246 Hughes, N. J., Chalk, P. A., Clayton, C. L., and Kelly, D. J. (1995) Identification of carboxylation enzymes and characterization of a novel foursubunit pyruvate:flavodoxin oxidoreductase from Helicobacter pylori. J. Bacteriol. 177, 39533959 Hughes, N. J., Clayton, C. L., Chalk, P. A., and Kelly, D. J. (1998) Helicobacter pylori porCDAB and oorDABC genes encode distinct pyruvate: flavodoxin and 2-oxoglutarate:acceptor oxidoreductases which mediate electron transport to NADP. J. Bacteriol. 1119 1128 Hoffman, P. S., Goodwin, A., Johnsen, J., Magee, K., and Veldhuyzen van Zanten, S. J. O. (1996) Metabolic activities of metronidazole-sensitive and -resistant strains of Helicobacter pylori: repression of pyruvate oxidoreductase and expression of isocitrate lyase activity correlate with resistance. J. Bacteriol. 178, 4822 4829 Ge, Z. M., Feng, Y., Dangler, C. A., Xu, S. L., Taylor, N. S., and Fox, J. G. (2000) Fumarate reductase is essential for Helicobacter pylori colonization of the mouse stomach. Microb. Pathog. 29, 279 287 Ge, Z. (2002) Potential of fumarate reductase as a novel therapeutic target in Helicobacter pylori infection. Expert Opin. Ther. Targets 6, 135146 Marshall, B. J., Barret, L., Prakash, C., McCallum, R., and Guerrant, R. (1990) Urea protects Helicobacter (Campylobacter) pylori from the bactericidal effect of acid. Gastroenterology 99, 269 276 Karita, M., Tsuda, M., and Nakazawa, T. (1995) Essential role of urease in vitro and in vivo Helicobacter pylori colonization study using a wild-type and isogenic urease mutant strain. J. Clin. Gastroenterol. 21, Suppl. 1, S160 S163 Backert, S., Gressmann, H., Kwok, T., Zimny Arndt, U., Konig, W., Jungblut, P. R., and Meyer, T. F. (2005) Gene expression and protein profiling of AGS gastric epithelial cells upon infection with Helicobacter pylori. Proteomics 5, 39023918 Ha, N. C., Oh, S. T., Sung, J. Y., Cha, K. A., Lee, M. H., and Oh, B. H. (2001) Supramolecular assembly and acid resistance of Helicobacter pylori urease. Nat. Struct. Biol. 8, 505509 Slonczewski, J. L., and Kirkpatrick, C. (2002) Proteomic analysis of pHdependent stress responses in Escherichia coli and Helicobacter pylori using two-dimensional gel electrophoresis. Methods Enzymol. 358, 228 242 Dunn, B. E., Vakil, N. B., Schneider, B. G., Miller, M. M., Zitzer, J. B., Peutz, T., and Phadnis, S. H. (1997) Localization of Helicobacter pylori urease and heat shock protein in human gastric biopsies. Infect. Immun. 65, 11811188

Molecular & Cellular Proteomics 6.2

205

Helicobacter pylori Proteomics

72. Censini, S., Lange, C., Xiang, Z., Crabtree, J. E., Ghiara, P., Borodovsky, M., Rappuoli, R., and Covacci, A. (1996) cag, a pathogenicity island of Helicobacter pylori, encodes type I-specific and disease-associated virulence factors. Proc. Natl. Acad. Sci. U. S. A. 93, 14648 14653 73. Odenbreit, S., Puls, J., Sedlmaier, B., Gerland, E., Fischer, W., and Haas, R. (2000) Translocation of Helicobacter pylori CagA into gastric epithelial cells by type IV secretion. Science 287, 14971500 74. Tanaka, J., Suzuki, T., Mimuro, H., and Sasakawa, C. (2003) Structural definition on the surface of Helicobacter pylori type IV secretion apparatus. Cell. Microbiol. 5, 395 404 75. van Doorn, L. J., Quint, W., Schneeberger, P., Tytgat, G. M., and de Boer, W. A. (1997) The only good Helicobacter pylori is a dead Helicobacter pylori. Lancet 350, 7172 76. Sprinzak, E., Sattath, S., and Margalit, H. (2003) How reliable are experimental protein-protein interaction data? J. Mol. Biol. 327, 919 923 77. Fischer, W., Buhrdorf, R., Gerland, E., and Haas, R. (2001) Outer membrane targeting of passenger proteins by the vacuolating cytotoxin autotransporter of Helicobacter pylori. Infect. Immun. 69, 6769 6775 78. Bumann, D., Aksu, S., Wendland, M., Janek, K., Zimny-Arndt, U., Sabarth, N., Meyer, T. F., and Jungblut, P. R. (2002) Proteome analysis of secreted proteins of the gastric pathogen Helicobacter pylori. Infect. Immun. 70, 3396 3403 79. Geis, G., Suerbaum, S., Forsthoff, B., Leying, H., and Opferkuch, W. (1993) Ultrastructure and biochemical studies of the flagellar sheath of Helicobacter pylori. J. Med. Microbiol. 38, 371377 80. Sanchez, S., Abel, A., Arenas, J., Criado, M. T., and Ferreiros, C. M. (2006) Cross-linking analysis of antigenic outer membrane protein complexes of Neisseria meningitidis. Res. Microbiol. 157, 136 142 81. Ferrero, R. L., Hazell, S. L., and Lee, A. (1988) The urease enzymes of Campylobacter pylori and a related bacterium. J. Med. Microbiol. 27, 33 40 82. Voland, P., Weeks, D. L., Marcus, E. A., Prinz, C., Sachs, G., and Scott, D. (2003) Interactions among the seven Helicobacter pylori proteins encoded by the urease gene cluster. Am. J. Physiol. 284, G96 G106 83. Sharp, J. A., and Edwards, M. R. (1978) Purification and properties of succinyl-coenzyme A-3-oxo acid coenzyme A-transferase from sheep kidney. Biochem. J. 173, 759 765 84. Jacob, U., Mack, M., Clausen, T., Huber, R., Buckel, W., and Messerschmidt, A. (1997) Glutaconate CoA-transferase from Acidaminococcus fermentans: the crystal structure reveals homology with other CoAtransferases. Structure (Lond.) 5, 415 426 85. Ge, Z. M., Jiang, Q., Kalisiak, M. S., and Taylor, D. E. (1997) Cloning and functional characterization of Helicobacter pylori fumarate reductase operon comprising three structural genes coding for subunits C, A and B. Gene (Amst.) 204, 227234 86. Mileni, M., MacMillan, F., Tziatzios, C., Zwicker, K., Haas, A. H., Mantele, W., Simon, J., and Lancaster, C. R. (2006) Heterologous production in Wolinella succinogenes and characterization of the quinol:fumarate reductase enzymes from Helicobacter pylori and Campylobacter jejuni. Biochem. J. 395, 191201

87. Unden, G., and Kroger, A. (1981) The function of the subunits of the fumarate reductase complex of Vibrio succinogenes. Eur. J. Biochem. 120, 577584 88. Lancaster, C. R., Kroger, A., Auer, M., and Michel, H. (1999) Structure of fumarate reductase from Wolinella succinogenes at 2.2 resolution. Nature 402, 377385 89. Kitano, K., Niimura, Y., Nishiyama, Y., and Miki, K. (1999) Stimulation of peroxidase activity by decamerization related to ionic strength: AhpC protein from Amphibacillus xylanus. J. Biochem. (Tokyo) 126, 313319 90. Wood, Z. A., Poole, L. B., Hantgan, R. R., and Karplus, P. A. (2002) Dimers to doughnuts: redox-sensitive oligomerization of 2-cysteine peroxiredoxins. Biochemistry 41, 54935504 91. Frazier, B. A., Pgeifer, J. D., Russel, D. G., Klak, P., Olsen, A. N., Hammar, M., Westblom, T. U., and Normark, S. T. (1993) Paracrystalline inclusions of a novel ferritin containing non-heme iron, produced by the human gastric pathogen Helicobacter pylori: evidence for a third class of ferritins. J. Bacteriol. 175, 966 972 92. Yariv, J., Kalb, A. J., Sperling, R., Bauminger, E. R., Cohen, S. G., and Ofer, S. (1981) The composition and the structure of bacterioferritin of Escherichia coli. Biochem. J. 197, 171175 93. Theil, E. C. (1987) Ferritin: structure, gene regulation, and cellular function in animals, plants, and microorganisms. Annu. Rev. Biochem. 56, 289 315 94. Spiegelhalder, C., Gerstenecker, B., Kersten, A., Schiltz, E., and Kist, M. (1993) Purification of Helicobacter pylori superoxide dismutase and cloning and sequencing of the gene. Infect. Immun. 61, 53155325 95. Beyer, W. F., Jr., Reynolds, J. A., and Fridovich, I. (1989) Differences between the manganese- and the iron-containing superoxide dismutases of Escherichia coli detected through sedimentation equilibrium, hydrodynamic, and spectroscopic studies. Biochemistry 28, 4403 4409 96. Bottomley, J. R., Clayton, C. L., Chalk, P. A., and Kleanthous, C. (1996) Cloning, sequencing, expression, purification and preliminary characterization of a type II dehydroquinase from Helicobacter pylori. Biochem. J. 319, 559 565 97. Lee, B. I., Kwak, J. E., and Suh, S. W. (2003) Crystal structure of the type II 3-dehydroquinase from Helicobacter pylori. Proteins 51, 616 617 98. Kinghorn, J. R., Schweizer, M., Giles, N. H., and Kushner, S. R. (1981) The cloning and analysis of the aroD gene of E. coli K-12. Gene (Amst.) 14, 73 80 99. Kleanthous, C., Deka, R., Davis, K., Kelly, S. M., Cooper, A., Harding, S. E., Price, N. C., Hawkins, A. R., and Coggins, J. R. (1992) A comparison of the enzymological and biophysical properties of two distinct classes of dehydroquinase enzymes. Biochem. J. 282, 687 695 100. Kimber, M. S., Martin, F., Lu, Y., Houston, S., Vedadi, M., Dharamsi, A., Fiebig, K. M., Schmid, M., and Rock, C. O. (2004) The structure of (3R)-hydroxyacyl-acyl carrier protein dehydratase (FabZ) from Pseudomonas aeruginosa. J. Biol. Chem. 279, 5259352602 101. Boneca, I. G., de Reuse, H., Epinat, J.-C., Pupin, M., Labigne, A., and Moszer, I. (2003) A revised annotation and comparative analysis of Helicobacter pylori genomes. Nucleic Acids Res. 31, 1704 1714

206

Molecular & Cellular Proteomics 6.2

You might also like